More stories

  • in

    Altered fire regimes modify lizard communities in globally endangered Araucaria forests of the southern Andes

    1.Krebs, P., Pezzatti, G. B., Mazzoleni, S., Talbot, L. M. & Conedera, M. Fire regime: History and definition of a key concept in disturbance ecology. Theory Biosci. 129, 53–69 (2010).PubMed 

    Google Scholar 
    2.Harvey, B. J., Donato, D. C. & Turner, M. G. Burn me twice, shame on who? Interactions between successive forest fires across a temperate mountain region. Ecology 97, 2272–2282 (2016).PubMed 

    Google Scholar 
    3.Prichard, S. J., Stevens-Rumann, C. S. & Hessburg, P. F. Tamm Review: Shifting global fire regimes: Lessons from reburns and research needs. For. Ecol. Manag. 396, 217–233 (2017).
    Google Scholar 
    4.González, M. E., Lara, A., Urrutia, R. & Bosnich, J. Cambio climático y su impacto potencial en la ocurrencia de incendios forestales en la zona centro-sur de Chile (33°–42° S). Bosque 32, 215–219 (2011).
    Google Scholar 
    5.Perfetti-Bolaño, A., González-acuña, D., Barrientos, C. & Moreno, L. Efectos del fuego sobre la avifauna del cerro Cayumanque, región del Bío-bío, Chile. Boletín Chil. Ornitol. 19, 1–11 (2013).
    Google Scholar 
    6.Engstrom, R. T. First-order fire effects on animals: Review and recommendations. Fire Ecol. 6, 115–130 (2010).
    Google Scholar 
    7.Doherty, T. S. et al. Ecosystem responses to fire: Identifying cross-taxa contrasts and complementarities to inform management strategies. Ecosystems 20, 872–884 (2017).
    Google Scholar 
    8.Kowaljow, E. et al. A 55-year-old natural experiment gives evidence of the effects of changes in fire frequency on ecosystem properties in a seasonal subtropical dry forest. Land Degrad. Dev. 30, 266–277 (2019).
    Google Scholar 
    9.Ferreira, C. C., Santos, X. & Carretero, M. A. Does ecophysiology mediate reptile responses to fire regimes? Evidence from Iberian lizards. PeerJ 4, e2107 (2016).PubMed 
    PubMed Central 

    Google Scholar 
    10.Russell, K. R., Van Lear, D. H. & Guynn, D. C. Prescribed fire effects on herpetofauna: Review and management implications. Wildl. Soc. Bull. 27, 374–384 (1999).
    Google Scholar 
    11.Shine, R., Brown, G. P. & Elphick, M. J. Effects of intense wildfires on the nesting ecology of oviparous montane lizards. Austral. Ecol. 41, 756–767 (2016).
    Google Scholar 
    12.Driscoll, D. A., Smith, A. L., Blight, S. & Maindonald, J. Reptile responses to fire and the risk of post-disturbance sampling bias. Biodivers. Conserv. 21, 1607–1625 (2012).
    Google Scholar 
    13.Hu, Y., Kelly, L. T., Gillespie, G. R. & Jessop, T. S. Lizard responses to forest fire and timber harvesting: Complementary insights from species and community approaches. For. Ecol. Manag. 379, 206–215 (2016).
    Google Scholar 
    14.Hromada, S. J. et al. Response of reptile and amphibian communities to the reintroduction of fire in an oak/hickory forest. For. Ecol. Manag. 428, 1–13 (2018).
    Google Scholar 
    15.Chergui, B., Pleguezuelos, J. M., Fahd, S. & Santos, X. Modelling functional response of reptiles to fire in two Mediterranean forest types. Sci. Total Environ. 732, 139205 (2020).ADS 
    CAS 
    PubMed 

    Google Scholar 
    16.Costa, B. M., Pantoja, D. L., Sousa, H. C., de Queiroz, T. A. & Colli, G. R. Long-term, fire-induced changes in habitat structure and microclimate affect Cerrado lizard communities. Biodivers. Conserv. 29, 1659–1681 (2020).
    Google Scholar 
    17.Gómez-González, S., Ojeda, F. & Fernandes, P. M. Portugal and Chile: Longing for sustainable forestry while rising from the ashes. Environ. Sci. Policy 81, 104–107 (2018).
    Google Scholar 
    18.Arroyo, M. T. K., Cavieres, L., Peñaloza, A., Riveros, M. & Faggi, A. Relaciones fitogeográficas y patrones regionales de riqueza de especies en la flora del bosque lluvioso templado de Sudamérica. In: Ecología de los Bosques Nativos de Chile (eds Armesto, J. et al.) 71–100 (1995).19.González, M. E., Veblen, T. T. & Sibold, J. S. Fire history of Araucaria-Nothofagus forests in Villarrica National Park, Chile. J. Biogeogr. 32, 1187–1202 (2005).
    Google Scholar 
    20.Veblen, T. T. Regeneration patterns in Araucaria araucana forests in Chile. J. Biogeogr. 9, 11 (1982).
    Google Scholar 
    21.Aagesen, D. L. Indigenous resource rights and conservation of the monkey-puzzle tree (Araucaria araucana, Araucariaceae): A case study from southern Chile. Econ. Bot. 52, 146–160 (1998).
    Google Scholar 
    22.Aagesen, D. Burning monkey-puzzle: Native fire ecology and forest management in northern Patagonia. Agric. Human Values 21, 233–242 (2004).
    Google Scholar 
    23.Pollmann, W. & Veblen, T. T. Nothofagus regeneration dynamics in south-central Chile: A test of a general model. Ecol. Monogr. 74, 615–634 (2004).
    Google Scholar 
    24.Ortega, M., Ponce, X. & Tamarín, R. Manual con medidas para la prevención de incendios forestales, IX Región (Corporación Nacional Forestal (CONAF), 2006).
    Google Scholar 
    25.Ferreira, D., Pinho, C., Brito, J. C. & Santos, X. Increase of genetic diversity indicates ecological opportunities in recurrent-fire landscapes for wall lizards. Sci. Rep. 9, 1–11 (2019).
    Google Scholar 
    26.Nimmo, D. G. et al. Predicting the century-long post-fire responses of reptiles. Glob. Ecol. Biogeogr. 21, 1062–1073 (2012).
    Google Scholar 
    27.Smith, A. L., Michael Bull, C. & Driscoll, D. A. Successional specialization in a reptile community cautions against widespread planned burning and complete fire suppression. J. Appl. Ecol. 50, 1178–1186 (2013).
    Google Scholar 
    28.Kelly, L. T., Bennett, A. F., Clarke, M. F. & Mccarthy, M. A. Optimal fire histories for biodiversity conservation. Conserv. Biol. 29, 473–481 (2015).PubMed 

    Google Scholar 
    29.Valentine, L. E., Reaveley, A., Johnson, B., Fisher, R. & Wilson, B. A. Burning in banksia woodlands: How does the fire-free period influence reptile communities?. PLoS ONE 7, e34448 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    30.Uribe, S. & Estades, C. F. Reptiles in monterey pine plantations of the coastal range of central Chile. Rev. Chil. Hist. Nat. 87, 1–8 (2014).
    Google Scholar 
    31.Santos, X., Badiane, A. & Matos, C. Contrasts in short- and long-term responses of Mediterranean reptile species to fire and habitat structure. Oecologia 180, 205–216 (2016).ADS 
    PubMed 

    Google Scholar 
    32.Ferreira, D., Mateus, C. & Santos, X. Responses of reptiles to fire in transition zones are mediated by bioregion affinity of species. Biodivers. Conserv. 25, 1543–1557 (2016).
    Google Scholar 
    33.Zúñiga, A. H. Changes in the structure of assemblages of three liolaemus lizards (Iguania, liolaemidae) in a protected area of south-central Chile affected by a mixed-severity wildfire. Zoodiversity 54, 265–274 (2020).
    Google Scholar 
    34.Rubio, A. V. & Simonetti, J. A. Lizard assemblages in a fragmented landscape of central Chile. Eur. J. Wildl. Res. 57, 195–199 (2011).
    Google Scholar 
    35.Driscoll, D. A. & Henderson, M. K. How many common reptile species are fire specialists? A replicated natural experiment highlights the predictive weakness of a fire succession model. Biol. Conserv. 141, 460–471 (2008).
    Google Scholar 
    36.Lindenmayer, D. B., Claridge, A. W., Gilmore, A. M., Michael, D. & Lindenmayer, B. D. The ecological roles of logs in Australian forests and the potential impacts of harvesting intensification on log-using biota. Pacific Conserv. Biol. 8, 121–140 (2002).
    Google Scholar 
    37.Evans, M. J., Newport, J. S. & Manning, A. D. A long-term experiment reveals strategies for the ecological restoration of reptiles in scattered tree landscapes. Biodivers. Conserv. 28, 2825–2843 (2019).
    Google Scholar 
    38.Mella, J. E. Guía de Campo Reptiles de Chile. Tomo: 1 Zona Central (2017).39.Whitford, K. R. & McCaw, W. L. Coarse woody debris is affected by the frequency and intensity of historical harvesting and fire in an open eucalypt forest. Aust. For. 82, 56–69 (2019).
    Google Scholar 
    40.Vidal, M. A. & Labra, A. Herpetología de Chile (GráficAndes, 2008).
    Google Scholar 
    41.Meiri, S. & Chapple, D. G. Biases in the current knowledge of threat status in lizards, and bridging the ‘assessment gap’. Biol. Conserv. 204, 6–15 (2016).
    Google Scholar 
    42.Tingley, R., Meiri, S. & Chapple, D. G. Addressing knowledge gaps in reptile conservation. Biol. Conserv. 204, 1–5 (2016).
    Google Scholar 
    43.Watson, J. E. M., Whittaker, R. J. & Dawson, T. P. Habitat structure and proximity to forest edge affect the abundance and distribution of forest-dependent birds in tropical coastal forests of southeastern Madagascar. Biol. Conserv. 120, 311–327 (2004).
    Google Scholar 
    44.Scott, D. M. et al. The impacts of forest clearance on lizard, small mammal and bird communities in the arid spiny forest, southern Madagascar. Biol. Conserv. 127, 72–87 (2006).
    Google Scholar 
    45.Bowman, D. M. J. S. et al. Vegetation fires in the Anthropocene. Nat. Rev. Earth Environ. 1, 500–515 (2020).ADS 

    Google Scholar 
    46.Hu, Y., Urlus, J., Gillespie, G., Letnic, M. & Jessop, T. S. Evaluating the role of fire disturbance in structuring small reptile communities in temperate forests. Biodivers. Conserv. 22, 1949–1963 (2013).
    Google Scholar 
    47.Gutiérrez, J. A., Krenz, J. D. & Ibargüengoytía, N. R. Effect of altitude on thermal responses of Liolaemus pictus argentinus in Argentina. J. Therm. Biol. 35, 332–337 (2010).
    Google Scholar 
    48.Artacho, P., Saravia, J., Perret, S., Bartheld, J. L. & Le Galliard, J. F. Geographic variation and acclimation effects on thermoregulation behavior in the widespread lizard Liolaemus pictus. J. Therm. Biol. 63, 78–87 (2017).PubMed 

    Google Scholar 
    49.Elzer, A. L. et al. Forest-fire regimes affect thermoregulatory opportunities for terrestrial ectotherms. Austral. Ecol. 38, 190–198 (2013).
    Google Scholar 
    50.Todd, B. D. & Andrews, K. M. Response of a reptile guild to forest harvesting. Conserv. Biol. 22, 753–761 (2008).PubMed 

    Google Scholar 
    51.Santos, X., Sillero, N., Poitevin, F. & Cheylan, M. Realized niche modelling uncovers contrasting responses to fire according to species-specific biogeographical affinities of amphibian and reptile species. Biol. J. Linn. Soc. 126, 55–67 (2019).
    Google Scholar 
    52.Farnsworth, L. M., Nimmo, D. G., Kelly, L. T., Bennett, A. F. & Clarke, M. F. Does pyrodiversity beget alpha, beta or gamma diversity? A case study using reptiles from semi-arid Australia. Divers. Distrib. 20, 663–673 (2014).
    Google Scholar 
    53.Vera-Escalona, I. M., Coronado, T., Muñoz-Mendoza, C. & Victoriano, P. F. Distribución histórica y actual de la lagartija Liolaemus pictus (Dumeril & Bibron 1837) (Liolaemidae) y nuevo límite continental sur de distribución. Gayana 74, 139–146 (2010).
    Google Scholar 
    54.Gunderson, A. R., Mahler, D. L. & Leal, M. Thermal niche evolution across replicated Anolis lizard adaptive radiations. Proc. R. Soc. B Biol. Sci. 285, 20172241 (2018).
    Google Scholar 
    55.Bowman, D. M. J. S. & Haberle, S. G. Paradise burnt: How colonizing humans transform landscapes with fire. Proc. Natl. Acad. Sci. USA. 107, 21234–21235 (2010).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    56.Maia-Carneiro, T., Dorigo, T. A. & Rocha, C. F. D. Influences of seasonality, thermal environment and wind intensity on the thermal ecology of Brazilian sand lizards in a restinga remnant. South Am. J. Herpetol. 7, 241–251 (2012).
    Google Scholar 
    57.Nimmo, D. G., Kelly, L. T., Farnsworth, L. M., Watson, S. J. & Bennett, A. F. Why do some species have geographically varying responses to fire history?. Ecography 37, 805–813 (2014).
    Google Scholar 
    58.Jones, G. M. et al. Megafires: An emerging threat to old-forest species. Front. Ecol. Environ. 14, 300–306 (2016).
    Google Scholar 
    59.Chergui, B., Fahd, S., Santos, X. & Pausas, J. G. Socioeconomic factors drive fire-regime variability in the mediterranean basin. Ecosystems 21, 619–628 (2018).
    Google Scholar 
    60.Hellmich, W. & Goetsch, W. Die eidechsen Chiles, insbesondere die gattung Liolaemus, nach den sammlungen Goetsch-Hellmich, Vol. 24 (1934).61.Veblen, T. T., Burns, B. R., Kitzberger, T., Lara, A., Villalba,
    A. The ecology of the conifers of southern South America. in Ecology of the Southern Conifers (eds Enright, N. J. & Hill, R. S.) 129–135 (Melbourne University Press, Carlton, Victoria, 1995).
    Google Scholar 
    62.Donoso, C. Bosques templados de Chile y Argentina. Variación, Estructura y Dinámica (Editorial Universitaria S.A., 1993).
    Google Scholar 
    63.Fuentes-Ramirez, A., Barrientos, M., Almonacid, L., Arriagada-Escamilla, C. & Salas-Eljatib, C. Short-term response of soil microorganisms, nutrients and plant recovery in fire-affected Araucaria araucana forests. Appl. Soil Ecol. 131, 99–106 (2018).
    Google Scholar 
    64.Urrutia-Estrada, J., Fuentes-Ramírez, A. & Hauenstein, E. Diferencias en la composición florística en bosques de Araucaria-Nothofagus afectados por distintas severidades de fuego. Gayana Bot. 75, 12–25 (2018).
    Google Scholar 
    65.González, M. E., Szejner, M., Muñoz, A. A. & Silva, J. Incendios catastróficos en bosques andinos de Araucaria-Nothofagus: Efecto de la severidad y respuesta de la vegetación. Bosque Nativ. 46, 12–17 (2009).
    Google Scholar 
    66.Luebert, F. & Pliscoff, P. Sinopsis bioclimática y vegetacional de Chile (Editorial Universitaria S.A., 2006).
    Google Scholar 
    67.(CONAF), C. N. F. Análisis de la afectación y severidad de los incendios forestales (2017).68.Zúñiga, A. H. et al. Rodent assemblage composition as indicator of fire severity in a protected area of south-central Chile. Austral. Ecol. 46, 249–260 (2021).
    Google Scholar 
    69.Demangel, D. Reptiles en Chile (Fauna Nativa Ediciones, 2016).
    Google Scholar 
    70.Vera-Escalona, I. et al. Lizards on ice: Evidence for multiple refugia in Liolaemus pictus (Liolaemidae) during the last glacial maximum in the southern Andean beech forests. PLoS ONE 7, e48358 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    71.Ibarra, J. T. & Martin, K. Biotic homogenization: Loss of avian functional richness and habitat specialists in disturbed Andean temperate forests. Biol. Conserv. 192, 418–427 (2015).
    Google Scholar 
    72.Buckland, S. T., Rexstad, E. A., Marques, T. A. & Oedekoven, C. S. Methods in Statistical Ecology (Springer, 2015).MATH 

    Google Scholar 
    73.Ibarra, J. T. & Martin, K. Beyond species richness: An empirical test of top predators as surrogates for functional diversity and endemism. Ecosphere 6, 1–15 (2015).
    Google Scholar 
    74.Royle, J. A., Dawson, D. K. & Bates, S. Modeling abundance effects in distance sampling. Ecology 85, 1591–1597 (2004).
    Google Scholar 
    75.Marques, T. A., Thomas, L., Fancy, S. G. & Buckland, S. T. Improving estimates of bird density using multiple-covariate distance sampling. Auk 124, 1229–1243 (2007).
    Google Scholar 
    76.Fiske, I. J. & Chandler, R. B. Unmarked: An R package for fitting hierarchical models of wildlife occurrence and abundance. J. Stat. Softw. 43, 1–23 (2011).
    Google Scholar 
    77.R Core Team. R: A Language and Environment for Statistical Computing (2019).78.Furnas, B. J., Newton, D. S., Capehart, G. D. & Barrows, C. W. Hierarchical distance sampling to estimate population sizes of common lizards across a desert ecoregion. Ecol. Evol. 9, 3046–3058 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    79.Burnham, K. P. & Anderson, D. R. Model Selection and Inference. A Practical Information-Theoretical Approach (Springer, 2002).MATH 

    Google Scholar 
    80.Pinheiro, J. & Bates, D. Package ‘nlme’: Linear and Nonlinear Mixed Effects Models (2020).81.Mazerolle, J. M. Package ‘AICcmodavg’: Model Selection and Multimodel Inference Based on (Q)AIC(c) (2020).82.Hothorn, T., Bretz, F. & Westfall, P. Simultaneous inference in general parametric models. Biometrical J. 50, 346–363 (2008).MathSciNet 
    MATH 

    Google Scholar  More

  • in

    Effects of soil texture and nitrogen fertilisation on soil bacterial community structure and nitrogen uptake in flue-cured tobacco

    Accumulation and distribution of N in flue-cured tobacco growing in different soilsAccumulation dynamics of N in different soilsNitrogen gradually increased in loam soil, clay loam, and sandy loam soils with plant growth (Fig. 1), attaining a maximum at the mature-plant stage(2.10 g/plant, 1.43 g/plant, and 2.90 g/plant, respectively). Nitrogen accumulation was lower in plants grown in clay loam than in plants grown in loam soil and sandy loam during the entire growth period, indicating that the N supply capacity of clay loam was relatively weak, and tobacco plants grown in this soil had the lowest levels of N uptake and utilisation. The N uptake and accumulation in flue-cured tobacco grown in loam soil and sandy loam were basically the same before the ceiling stage, but at the mature stage, N accumulation was significantly higher in plants grown in sandy loam than in plants grown in loam soil and clay loam (P  More

  • in

    Inferring functional traits in a deep-sea wood-boring bivalve using dynamic energy budget theory

    1.Howell, K. L. et al. A decade to study deep-sea life. Nat. Ecol. Evol. https://doi.org/10.1038/s41559-020-01352-5 (2020).Article 

    Google Scholar 
    2.Howell, K. L. et al. A blueprint for an inclusive, global deep-sea ocean decade field program. Front. Mar. Sci. 7, 1–25. https://doi.org/10.3389/fmars.2020.584861 (2020).ADS 
    Article 

    Google Scholar 
    3.Ramirez-Llodra, E. et al. Man and the last great wilderness: Human impact on the deep sea. PLoS ONE 6, 22588. https://doi.org/10.1371/journal.pone.0022588 (2011).ADS 
    CAS 
    Article 

    Google Scholar 
    4.Bernardino, A. F., Levin, L. A., Thurber, A. R. & Smith, C. R. Comparative composition, diversity and trophic ecology of sediment macrofauna at vents, seeps and organic falls. PLoS ONE 7, e33515. https://doi.org/10.1371/journal.pone.0033515 (2012).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    5.Thiel, M. & Gutow, L. The ecology of rafting in the marine environment. II. The rafting organisms and community. Ocean. Mar. Biol. 43, 279–418. https://doi.org/10.1201/9781420037449.ch7 (2005).Article 

    Google Scholar 
    6.McClain, C. & Barry, J. Beta-diversity on deep-sea wood falls reflects gradients in energy availability. Biol. Lett. 10, 20140129. https://doi.org/10.1098/rsbl.2014.0129 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    7.Knudsen, J. The Bathyal and Abyssal Xylophaga (Pholadidae, Bivalvia) (Danish Science Press Ltd., 1961).
    Google Scholar 
    8.Turner, R. Wood-boring bivalves, opportunistic species in the deep sea. Science 180, 1377–1379. https://doi.org/10.1126/science.180.4093.1377 (1973).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    9.Voight, J. R. Deep-sea wood-boring bivalves of Xylophaga (Myoida: Pholadidae) on the continental shelf: A new species described. J. Mar. Biol. Assoc. UK 88, 1459–1464. https://doi.org/10.1017/S0025315408002117 (2008).Article 

    Google Scholar 
    10.Turner, R. D. A survey and Illustrated Catalogue of the Teredinidae (Mollusca: Bivalvia) (Harvard University, 1966).Book 

    Google Scholar 
    11.Hoppe, K. N. Teredo Navalis—the Cryptogenic Shipworm. in Invasive Aquatic Species of Europe. Distribution, Impacts and Management. (ed. Leppäkoski E., Gollasch S., O. S.) 116–119, https://doi.org/10.1007/978-94-015-9956-6_12 (2002).12.Distel, D. L. & Roberts, S. J. Bacterial endosymbionts in the gills of the deep-sea wood-boring bivalves Xylophaga atlantica and X. washingtona. Biol. Bull. 192, 253–261. https://doi.org/10.2307/1542719 (1997).CAS 
    Article 
    PubMed 

    Google Scholar 
    13.Distel, D. L., Morrill, W., MacLaren-Toussaint, N., Franks, D. & Waterbury, J. Teredinibacter turnerae gen. nov., sp. Nov., a dinitrogen-fixing, cellulolytic, endosymbiotic gamma-proteobacterium isolated from the gills of wood-boring molluscs (Bivalvia: Teredinidae). Int. J. Syst. Evol. Microbiol. 52, 2261–2269 (2002).CAS 
    PubMed 

    Google Scholar 
    14.O’Connor, R. M. et al. Gill bacteria enable a novel digestive strategy in a wood-feeding mollusk. Proc. Natl. Acad. Sci. U. S. A. 111, 5096–5104. https://doi.org/10.1073/pnas.1413110111 (2014).CAS 
    Article 

    Google Scholar 
    15.Sabbadin, F. et al. Uncovering the molecular mechanisms of lignocellulose digestion in shipworms. Biotechnol. Biofuels 11, 1–14. https://doi.org/10.1186/s13068-018-1058-3 (2018).CAS 
    Article 

    Google Scholar 
    16.Kooijman, S. A. L. M. Dynamic Energy Budget Theory for Metabolic Organisation (Cambridge University Press, 2010).
    Google Scholar 
    17.Sarà, G., Palmeri, V., Montalto, V., Rinaldi, A. & Widdows, J. Parameterisation of bivalve functional traits for mechanistic eco-physiological dynamic energy budget (DEB) models. Mar. Ecol. Prog. Ser. 480, 99–117. https://doi.org/10.3354/meps10195 (2013).ADS 
    Article 

    Google Scholar 
    18.Sarà, G., Rinaldi, A. & Montalto, V. Thinking beyond organism energy use: A trait-based bioenergetic mechanistic approach for predictions of life-history traits in marine organisms. Mar. Ecol. 35, 506–515. https://doi.org/10.1111/maec.12106 (2014).ADS 
    Article 

    Google Scholar 
    19.Mangano, M. C. et al. Moving toward a strategy for addressing climate displacement of marine resources: A proof-of-concept. Front. Mar. Sci. 7, 1–16. https://doi.org/10.3389/fmars.2020.00408 (2020).ADS 
    Article 

    Google Scholar 
    20.Romano, C. et al. Wooden stepping stones: Diversity and biogeography of deep-sea wood-boring Xylophagaidae (Mollusca: Bivalvia) in the North-East Atlantic Ocean, with the description of a new genus. Front. Mar. Sci. https://doi.org/10.3389/fmars.2020.579959 (2020).Article 

    Google Scholar 
    21.Culliney, J. L. & Turner, R. D. Larval development of the deep-water wood boring bivalve, Xylophaga atlantica Richards (Mollusca, bivalvia, pholadidae). Ophelia 15, 149–161. https://doi.org/10.1080/00785326.1976.10425455 (1976).Article 

    Google Scholar 
    22.Romey, W., Bullock, R. & Dealteris, J. Rapid growth of a deep-sea wood-boring bivalve. Cont. Shelf Res. 14, 1349–1359. https://doi.org/10.1016/0278-4343(94)90052-3 (1994).ADS 
    Article 

    Google Scholar 
    23.Gaudron, S. M. et al. Colonization of organic substrates deployed in deep-sea reducing habitats by symbiotic species and associated fauna. Mar. Environ. Res. 70, 1–12. https://doi.org/10.1016/j.marenvres.2010.02.002 (2010).CAS 
    Article 
    PubMed 

    Google Scholar 
    24.Gaudron, S. M., Haga, T., Wang, H., Laming, S. R. & Duperron, S. Plasticity in reproduction and nutrition in wood-boring bivalves (Xylophaga atlantica) from the Mid-Atlantic Ridge. Mar. Biol. 163, 1–12. https://doi.org/10.1007/s00227-016-2988-6 (2016).CAS 
    Article 

    Google Scholar 
    25.Childress, J. J., Cowles, D. L., Favuzzi, J. A. & Mickel, T. J. Metabolic rates of benthic deep-sea decapod crustaceans decline with increasing depth primarily due to the decline in temperature. Deep Sea Res. Part A Oceanogr. Res. Pap. 37, 929–949. https://doi.org/10.1016/0198-0149(90)90104-4 (1990).ADS 
    CAS 
    Article 

    Google Scholar 
    26.Childress, J. J. Are there physiological and biochemical adaptations of metabolism in deep-sea animals?. Trends Ecol. Evol. 10, 30–36. https://doi.org/10.1016/S0169-5347(00)88957-0 (1995).CAS 
    Article 
    PubMed 

    Google Scholar 
    27.Tittensor, D. P., Rex, M. A., Stuart, C. T., Mcclain, C. R. & Smith, C. R. Species—energy relationships in deep-sea molluscs subject collections species—energy relationships in deep-sea molluscs. Biol. Lett. 7, 718–722 (2011).Article 

    Google Scholar 
    28.McClain, C. R., Allen, A. P., Tittensor, D. P. & Rex, M. A. Energetics of life on the deep seafloor. Proc. Natl. Acad. Sci. U. S. A. 109, 15366–15371. https://doi.org/10.1073/pnas.1208976109 (2012).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    29.Mickel, T. J. & Childress, J. J. Effects of pressure and temperature on the EKG and heart rate of the hydrothermal vent crab Bythograea Thermydron (Brachyura). Biol. Bull. 162, 70–82. https://doi.org/10.2307/1540971 (1982).Article 

    Google Scholar 
    30.Voight, J. R., Cooper, J. C. & Lee, R. W. Stable isotopic evidence of mixotrophy in Xylophagaids, deep-sea wood-boring bivalves. Front. Mar. Sci. 7, 50. https://doi.org/10.3389/fmars.2020.00050 (2020).Article 

    Google Scholar 
    31.Lika, K. et al. The ‘covariation method’ for estimating the parameters of the standard dynamic energy budget model I: Philosophy and approach. J. Sea Res. 66, 270–277. https://doi.org/10.1016/j.seares.2011.07.010 (2011).ADS 
    Article 

    Google Scholar 
    32.Marques, G. M. et al. The AmP project: Comparing species on the basis of dynamic energy budget parameters. PLoS Comput. Biol. 14, 1–23. https://doi.org/10.1371/journal.pcbi.1006100 (2018).CAS 
    Article 

    Google Scholar 
    33.Mariño, J., Augustine, S., Dufour, S. C. & Hurford, A. Dynamic Energy Budget theory predicts smaller energy reserves in thyasirid bivalves that harbour symbionts. J. Sea Res. 143, 119–127. https://doi.org/10.1016/j.seares.2018.07.015 (2019).ADS 
    Article 

    Google Scholar 
    34.Brown, A. et al. Metabolic costs imposed by hydrostatic pressure constrain bathymetric range in the lithodid crab Lithodes maja. J. Exp. Biol. 220, 3916–3926. https://doi.org/10.1242/jeb.158543 (2017).Article 
    PubMed 

    Google Scholar 
    35.Eisenmenger, M. J. & Reyes-De-Corcuera, J. I. High pressure enhancement of enzymes: A review. Enzyme Microb. Technol. 45, 331–347. https://doi.org/10.1016/j.enzmictec.2009.08.001 (2009).CAS 
    Article 

    Google Scholar 
    36.Kalenitchenko, D. et al. Bacteria alone establish the chemical basis of the wood-fall chemosynthetic ecosystem in the deep-sea. ISME J. 12, 367–379. https://doi.org/10.1038/ismej.2017.163 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    37.Levesque, C., Limén, H. & Juniper, S. K. Origin, composition and nutritional quality of particulate matter at deep-sea hydrothermal vents on Axial Volcano NE pacific. Mar. Ecol. Prog. Ser. 289, 43–52. https://doi.org/10.3354/meps289043 (2005).ADS 
    Article 

    Google Scholar 
    38.Limén, H., Levesque, C. & Kim Juniper, S. POM in macro-/meiofaunal food webs associated with three flow regimes at deep-sea hydrothermal vents on Axial Volcano, Juan de Fuca Ridge. Mar. Biol. 153, 129–139. https://doi.org/10.1007/s00227-007-0790-1 (2007).Article 

    Google Scholar 
    39.Culliney, J. L. Comparative larval development of the shipworms Bankia gouldi and Teredo navalis. Mar. Biol. 29, 245–251. https://doi.org/10.1007/BF00391850 (1975).Article 

    Google Scholar 
    40.Ramirez Llodra, E. Fecundity and life-history strategies in marine invertebrates. Adv. Mar. Biol. 43, 87–170. https://doi.org/10.1016/S0065-2881(02)43004-0 (2002).Article 
    PubMed 

    Google Scholar 
    41.Fernandez-Arcaya, U. et al. Bathymetric gradients of fecundity and egg size in fishes: A Mediterranean case study. Deep Sea Res. Part A Oceanogr. Res. Pap. 116, 106–117. https://doi.org/10.1016/j.enzmictec.2009.08.001 (2016).ADS 
    CAS 
    Article 

    Google Scholar 
    42.Young, C. M., Emson, R. H., Rice, M. E. & Tyler, P. A. A paradoxical mismatch of fecundity and recruitment in deep-sea opportunists: cocculinid and pseudococculinid limpets colonizing vascular plant remains on the Bahamian Slope. Deep Sea Res. 92, 36–45. https://doi.org/10.1016/j.dsr2.2013.01.027 (2013).ADS 
    Article 

    Google Scholar 
    43.Thorson, G. Reproductive and larval ecology of marine bottom invertebrates. Biol. Rev. 25, 1–45. https://doi.org/10.1111/j.1469-185X.1950.tb00585.x (1950).CAS 
    Article 
    PubMed 

    Google Scholar 
    44.Hitt, N. T. et al. Growth and longevity of New Zealand black corals. Deep. Res. Part I Oceanogr. Res. Pap. 162, e103298. https://doi.org/10.1016/j.dsr.2020.103298 (2020).Article 

    Google Scholar 
    45.McNichol, J. et al. Primary productivity below the seafloor at deep-sea hot springs. Proc. Natl. Acad. Sci. U. S. A. 115, 6756–6761. https://doi.org/10.1073/pnas.1804351115 (2018).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    46.Levin, L. A. et al. Hydrothermal vents and methane seeps: Rethinking the sphere of influence. Front. Mar. Sci. 3, 1–23. https://doi.org/10.3389/fmars.2016.00072 (2016).ADS 
    Article 

    Google Scholar 
    47.Nedoncelle, K., Lartaud, F., de Rafelis, M., Boulila, S. & Le Bris, N. A new method for high-resolution bivalve growth rate studies in hydrothermal environments. Mar. Biol. 160, 1427–1439. https://doi.org/10.1007/s00227-013-2195-7 (2013).CAS 
    Article 

    Google Scholar 
    48.Turekian, K. K., Cochran, J. K. & Bennett, J. T. Growth rate of a vesicomyid clam from the 21° N East Pacific Rise hydrothermal area. Nature 303, 55–56. https://doi.org/10.1038/303055a0 (1983).ADS 
    Article 

    Google Scholar 
    49.Lutz, R. A. et al. Rapid growth at deep-sea vents. Nature 371, 663–664. https://doi.org/10.1038/371663a0 (1994).ADS 
    Article 

    Google Scholar 
    50.Reed, A. J., Morris, J. P., Linse, K. & Thatje, S. Plasticity in shell morphology and growth among deep-sea protobranch bivalves of the genus Yoldiella (Yoldiidae) from contrasting Southern ocean regions. Deep. Res. Part I Oceanogr. Res. Pap. 81, 14–24. https://doi.org/10.1016/j.dsr.2013.07.006 (2013).ADS 
    Article 

    Google Scholar 
    51.Oliver, G., Allen, J. A. & Yonge, M. The functional and adaptive morphology of the deep-sea species of the Arcacea (Mollusca: Bivalvia) from the Atlantic. Philos. Trans. R. Soc. London. B Biol. Sci. 291, 45–76. https://doi.org/10.1098/rstb.1980.0127 (1980).ADS 
    Article 

    Google Scholar 
    52.Romano, C., Voight, J. R., Pérez-Portela, R. & Martin, D. Morphological and genetic diversity of the wood-boring Xylophaga (Mollusca, Bivalvia): New species and records from deep-sea Iberian canyons. PLoS ONE 9, 102887. https://doi.org/10.1371/journal.pone.0102887 (2014).ADS 
    CAS 
    Article 

    Google Scholar 
    53.Saulsbury, J. et al. Evaluating the influences of temperature, primary production, and evolutionary history on bivalve growth rates. Paleobiology 45, 405–420. https://doi.org/10.1017/pab.2019.20 (2019).Article 

    Google Scholar 
    54.Moss, D. K. et al. Lifespan, growth rate, and body size across latitude in marine bivalvia, with implications for phanerozoic evolution. Proc. R. Soc. B Biol. Sci. 283, 20161364. https://doi.org/10.1098/rspb.2016.1364 (2016).Article 

    Google Scholar 
    55.Tyler, P. A., Young, C. M. & Dove, F. Settlement, growth and reproduction in the deep-sea wood-boring bivalve mollusc Xylophaga depalmai. Mar. Ecol. Prog. Ser. 343, 151–159. https://doi.org/10.3354/meps06832 (2007).ADS 
    Article 

    Google Scholar 
    56.Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M. & West, G. B. Toward a metabolic theory of ecology. Ecology 85, 1771–1789. https://doi.org/10.1890/03-9000 (2004).Article 

    Google Scholar 
    57.Maino, J. L., Kearney, M. R., Nisbet, R. M. & Kooijman, S. A. L. M. Reconciling theories for metabolic scaling. J. Anim. Ecol. 83, 20–29. https://doi.org/10.1111/1365-2656.12085 (2014).Article 
    PubMed 

    Google Scholar 
    58.Gaudron, S. M., Demoyencourt, E. & Duperron, S. Reproductive traits of the cold-seep symbiotic mussel Idas modiolaeformis: gametogenesis and larval biology. Biol. Bull. 222, 6–16. https://doi.org/10.1086/bblv222n1p6 (2012).Article 

    Google Scholar 
    59.Hilário, A. et al. Estimating dispersal distance in the deep sea: Challenges and applications to marine reserves. Front. Mar. Sci. 2, 6. https://doi.org/10.3389/fmars.2015.00006 (2015).ADS 
    Article 

    Google Scholar 
    60.Marsh, A. G., Mullineaux, L. S., Young, C. M. & Manahan, D. T. Larval dispersal potential of the tubeworm Riftia pachyptila at deep-sea hydrothermal vents. Nature 411, 77–80. https://doi.org/10.1038/35075063 (2001).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    61.Young, C. M. et al. Dispersal of deep-sea larvae from the intra-American seas: Simulations of trajectories using ocean models. Integr. Comp. Biol. 52, 483–496. https://doi.org/10.1093/icb/ics090 (2012).Article 
    PubMed 

    Google Scholar 
    62.Yearsley, J. M., Salmanidou, D. M., Carlsson, J., Burns, D. & Van Dover, C. L. Biophysical models of persistent connectivity and barriers on the northern Mid-Atlantic Ridge. Deep. Res. Part II Top. Stud. Oceanogr. 180, 104819. https://doi.org/10.1016/j.dsr2.2020.104819 (2020).Article 

    Google Scholar 
    63.Levin, L. A. et al. Global observing needs in the deep ocean. Front. Mar. Sci. 6, 1–32. https://doi.org/10.3389/fmars.2019.00241 (2019).ADS 
    Article 

    Google Scholar 
    64.McClain, C. R., Boyer, A. G. & Rosenberg, G. The island rule and the evolution of body size in the deep sea. J. Biogeogr. 33, 1578–1584. https://doi.org/10.1111/j.1365-2699.2006.01545.x (2006).Article 

    Google Scholar 
    65.Zonneveld, C. & Kooijman, S. A. L. M. Application of a dynamic energy budget model to Lymnaea stagnalis (L.). Funct. Ecol. 3, 269–278. https://doi.org/10.2307/2389365 (1989).Article 

    Google Scholar 
    66.Mueller, C. A., Augustine, S., Kooijman, S. A. L. M., Kearney, M. R. & Seymour, R. S. The trade-off between maturation and growth during accelerated development in frogs. Comp. Biochem. Physiol. A 163, 95–102. https://doi.org/10.1016/j.cbpa.2012.05.190 (2012).CAS 
    Article 

    Google Scholar 
    67.MacArthur, R.H. & Wilson, E. The Theory of Island Biogeography (1967).68.Kooijman, S. A. L. M. Metabolic acceleration in animal ontogeny: An evolutionary perspective. J. Sea Res. 94, 128–137. https://doi.org/10.1016/j.seares.2014.06.005 (2014).ADS 
    Article 

    Google Scholar  More

  • in

    Highly efficient and simultaneous catalytic reduction of multiple toxic dyes and nitrophenols waste water using highly active bimetallic PdO–NiO nanocomposite

    The crystalline nature and phase purity of the synthesized NiO, PdO and PdO–NiO nanocomposite were confirmed by XRD analysis and shown in Fig. 1a. All three XRD patterns show the highly crystalline nature, which confirms the purity of the samples. In PdO–NiO mixed metal oxide, The intense diffraction peaks at 2θ = 37.16, 43.24, 62.81, 75.32 and 79.34° indexed to (101), (012), (110), (113) and (202) planes of the cubic phase of NiO, which is highly consistent with standard JCPDS NO: 01-071-475120. On the other hand, the sharp peaks at 2θ = 34.54 and 55.79° indexed to (101) and (112) planes corresponding to the tetragonal crystalline phase of PdO with an average lattice parameter 3.043 Å, which is highly associated with standard JCPDS NO: 043-102421. Interestingly, metallic diffraction peaks and other impurity phases were not observed in the hybrid PdO–NiO nanocomposite. This confirming that metal source (Pd) completely oxidized to the metal oxide (PdO) and formed hybrid PdO–NiO nanocomposite. In addition, the single metal oxide (PdO, NiO) diffraction patterns were compared in Fig. 1a.Figure 1(a) X-ray diffraction pattern and (b) FTIR spectra of NiO, PdO and PdO–NiO (c) TGA curve of PdO–NiO nanocomposite.Full size imageThe sharp peaks of each metal oxide confirm the high crystallinity of the tetragonal phase of the PdO and cubic phase of NiO. Both diffraction pattern of the single metal oxides is well associated with standard JCPDS file: 01-071-4751 and 043-1024 respectively. The average crystallite size, micro-strains of the NiO, PdO and PdO doped NiO were calculated by using Scherrer analysis and W–H analysis and plotted in Fig. 1b. A small variation was observed in the crystallite size of the catalysts (PdO, NiO and PdO–NiO) in Fig. 1b, which is due to the difference in the distribution of the crystal in the catalysts. The average crystallite size of the PdO, NiO and PdO–NiO was found to be 10.8, 7.8 and 7.36 nm, respectively. The PdO doping reduces the crystallite size in the PdO–NiO composite, consistent with previous reports. The calculated crystallite sizes, d-spacing, micro-strains, and binding energies of the PdO, NiO, and PdO–NiO are shown in the Table 1. In addition, the experimentally calculated d-spacing value of the pure PdO and NiO was well correlated with theoretical values and shown in Table SI. 1. The synthesized catalyst showed the negative and positive slopes of ε are corresponds to the compressive and tensile stress, respectively.Table 1 Crystallite sizes, d-spacing, micro-strains, and binding energies of the PdO, NiO, and PdO–NiO.Full size tableFurthermore, the chemical bonding and functional groups were analyzed by FTIR spectrometer. Figure 1c show FTIR spectra of the synthesized NiO, PdO and PdO–NiO nanocomposite. Similar spectra were observed for the three catalyst, the absorption peaks at high-frequency region 3200–3400 cm−1 belongs to O–H stretching vibration of the water molecules, due to surface adsorption phenomenon. Furthermore, three absorption peaks appeared at 1398.2, 1237.8 and 1057.8 cm−1, which ascribed to the C–O, CH2 and C=O stretching vibrations, which is well associated with XPS analysis data. The metal oxide bonding peaks appeared in the frequency range of 480.2–702.6 cm−1. Hence the Pd–O and Ni–O stretching frequency in the PdO doped NiO sample confirmed the formation of hybrid PdO–NiO nanocomposite22. After that, the thermal stability of the PdO–NiO nanocomposite was studied by TGA analysis. Figure 1d shows the thermogram of the PdO–NiO nanocomposite. The first weight loss (7%) started in the temperature range of 65 to 180 °C. Due to the H2O molecules, evaporation and then the sustainable weight loss of around 10% was observed in the range of 187 to 574 °C. Beyond 600 °C a significant weight loss 20% was observed, which may be assigned to the unreacted CO3 combustion23. Hence, the XRD, FTIR and TGA spectral studies confirmed the formation of hybrid PdO–NiO nanocomposite.The electronic state and chemical bonding of the PdO doped NiO composite was analyzed by using XPS spectra. Figure 2 shows the XPS spectra of the PdO–NiO nanocomposite, the broad scan spectrum (Fig. SI. 1) of the PdO–NiO, which show the existence of the Pd (3d), Ni (2p), O (1s) and C(1s) elements. The deconvoluted Pd 3d XPS spectrum in Fig. 2a shows two major peaks at a binding energy of 336.4 and 342.2 eV corresponds to spin–orbit doublets of Pd 3d5/2 and Pd 3d3/2, respectively, which confirmed Pd2+ ions in the form of PdO in the PdO–NiO nanocomposite24,25. In addition, the satellite peak of Pd species appeared at a binding energy of 339.2 eV and 345.3 eV. In Ni 2p spectra (Fig. 2b), Ni 2p3/2 and Ni 2p3/2 spin–orbit doublets peaks were observed at 854.9 eV and 872.5 eV, which corresponded to Ni–O and Ni–OH, respectively and their corresponding satellite peaks are located at a binding energy of 867.2 eV and 879.1 eV26,27. Furthermore, O 1s spectra (Fig. 2c) show the two peaks at a binding energy of 529.3 eV and 534.2 eV, which ascribed M–O and M–OH species. The obtained XPS spectra of the PdO–NiO nanocomposite are well associated with the XRD and EDX analysis.Figure 2High resolution X-ray photoelectron spectroscopy (a) Pd 3d, (b) Ni 2p, (c) O 1s and (d) C 1s spectra of PdO–NiO nanocomposite.Full size imageThe morphology feature and elemental composition of the prepared NiO, PdO and PdO–NiO nanocomposite was scrutinised by FE-SEM. Figure 3 shows the SEM morphology images of NiO, PdO and PdO–NiO nanocomposite at low and high magnification. Pure metal oxides NiO and PdO samples in Fig. 3a–d shows the porous crystalline morphology with high purity of the respective elements. On the other hand, PdO doped NiO sample in Fig. 3e,f shows the uniform, monodisperse, spherical crystalline morphology. Which confirms that PdO uniformly distributed with NiO. Hence, the PdO doping enhances the surface area of the catalyst. In addition, the elemental composition and elemental mapping was analysed for PdO–NiO sample and shown in Fig. 3g,h. The EDX spectra and elemental mapping clearly confirms the presence of the Pd, Ni and O elements in the composite with high purity.Figure 3FESEM images (a,b) NiO (c,d) PdO, (e,f) PdO–NiO and (g,f) EDS spectra and elemental mapping of PdO–NiO nanocomposite.Full size imageThe detailed morphology and particle size distribution of the PdO–NiO NPs was measured by HR-TEM and the results are presented in Fig. 4. Figure 4a–c shows the typical HRTEM images of the as-synthesized PdO–NiO nanocomposite. The obtained TEM images confirmed the uniform distribution of the spherical PdO–NiO NPs, which agrees with FESEM results. In Fig. 4c (inset), the histogram reveals that formed PdO–NiO nanoparticles are uniformly distributed with an average particle size of about 9.64 ± 2.1 nm, which is well associated with XRD crystallite size. Furthermore, the SAED pattern was analyzed to understand the crystallinity and the crystal quality of the PdO–NiO nanoparticles are shown in Fig. 4d. The clear ring-like structure suggests the polycrystalline nature of PdO–NiO. The obtained diffraction rings d-spacing values are corresponding to the (101), (012), (110), (113) and (202) planes of the NiO nanoparticles. Figure 4e shows the lattice fringes of the PdO doped NiO nanoparticles. The fringes show the lattice planes for both metal oxides. The interplanar d-spacing value of 0.1997 nm to correspond to the (012) plane of the NiO phase and the d-spacing value of 0. 2145 nm to correspond to the (110) plane of the PdO in the composite. Which is well correlated with the XRD d-spacing values. Elemental mapping in Fig. 4f shows the presence of Ni, Pd and O elements with uniform distribution as similar as SEM mapping. The morphology results of the synthesized catalysts are well associated with XRD, FTIR and XPS analysis.Figure 4High resolution transmission electron microscopy HRTEM images (a–c), (d) SAED pattern, (e) interplanr spacing and (f) elemental mapping of PdO–NiO nanocomposite.Full size imageUV–Vis absorption spectra were analyzed for the as-synthesized catalysts NiO, PdO and PdO–NiO and the respective results are presented in Fig. 5a. The absorption spectra show the strongest absorption maxima at 234.8 nm for all three catalysts. In addition, the characteristic absorption band of NiO and PdO were observed at 338.2 nm and 422.1 nm respectively23, on the other hand, no characteristic absorption band was observed for PdO–NiO sample. Furthermore, the bandgap energy was calculated for three catalysts by using the Schuster-Kubelka–Munk function.$$(alpha {text{h}}nu ),{text{n}} = {text{A}}({text{h}}nu – {text{Eg}})$$
    (1)
    Figure 5(a) UV–Vis absorption spectra and (b) Plot of (αhν)2 Vs hν for NiO, PdO and PdO–NiO nanocomposite (Inset shows the bandgap energy of the catalyst).Full size imageThe bandgap energy (Eg) was achieved by extrapolating against the photon energy and the obtained results are shown in Fig. 5b. The calculated bandgap (Eg) of NiO, PdO and PdO–NiO are 4.05 eV, 3.84 eV and 4.24 eV, respectively25 (Fig. 5b, inset). The PdO doping with NiO increases its bandgap value. This suggests that the PdO interface and NiO interface are closely combined in the composite. The obtained band gap value of the catalysts is much higher than the reported bandgap energy. The bandgap energy is highly dependent on the particle’s size. The bandgap energy increases with decreasing particle size, which confirmed that the synthesized catalysts are in nanoscale. The bandgap energy (Eg) of the PdO–NiO catalyst is well associated with FESEM and TEM results.The photoluminescence (Pl) spectra of NiO, PdO and PdO–NiO materials were measured at 325 nm excitation wavelength and presented in Fig. 6. Figure 6a shows the Pl spectra of the pure NiO, PdO and PdO–NiO nanocomposite. The blue/violet emission was observed for all three samples at 364 nm due to the excitation of 3d8 electrons of Ni2+ ions from the conduction band to the valence band24. From Fig. 6a, it can be seen that the intensity of the PdO–NiO nanocomposite is lower than the pure NiO and PdO, which indicated the higher electron transfer between the NiO and PdO, which is well correlated with electrochemical results. The deconvoluted PL spectra of NiO, PdO and PdO–NiO materials are shown in Fig. 6b–d; four peaks have been fitted for each sample as shown in Fig. 6b–d. The UV emission at 364 nm (3.4 eV) corresponds to the near band edge (NBE) excitation of NiO25. The obtained PL spectra confirmed that the PdO–NiO nanocomposite has more conductivity than the pure metal oxides.Figure 6(a) PL spectra and (b–d) deconvoluted spectra of NiO, PdO and PdO–NiO nanocomposite respectively.Full size imageThe electrode kinetics of the NiO, PdO and PdO–NiO modified GC electrode was explored in 1 M KOH at different scan rate variation at room temperature. In addition, the resistance of the aforesaid electrodes was monitored in impedance analysis and shown in Fig. 7. Figure 7a, the NiO/GC show a pair of well-defined redox peaks at around 0.49 V and 0.44 V respectively, which corresponding to the reversible reaction between Ni2+ and Ni3+26. In addition, the redox peak currents linearly increase with increasing scan rate.Figure 7(a–c) Cyclic voltammetry and (d) EIS curves of NiO, PdO and PdO–NiO nanocomposite in 1 M KOH electrolyte solution.Full size imageFigure 7b shows CV pattern of PdO/GC electrode in 1 M KOH solution, which sows poor peaks palladium oxide and palladium reduction peak at 0.52 V and 0.53 V respectively due to the formation of oxyhydroxide on the electrode surface in basic medium. Whereas the PdO–NiO NPs modified GC electrode in Fig. 7c show well-defined Ni2+ and Ni3+ kinetics with eightfold higher peak current. Which confirms the Efficient electron transfer between NiO and PdO in the composite. Which is well associated with PL results. Furthermore, the impedance spectra were achieved for three electrodes at fixed over potential (500 mV s−1) in 1 M KOH electrolyte and presented in Fig. 7d. In Fig. 7d the Rct value of the pure metal oxides NiO/GC and PdO/GC electrode were obtained as 236 Ω and 2702 Ω respectively. Whereas the bimetal oxide PdO–NiO/GC show 425.7 Ω, which is lower than the pure PdO. Due to the superior electron transformation between each metal oxide.Catalytic reduction of Azo compoundsAzo compounds are highly toxic to the environment as well as human beings. Especially, nitrophenols are listed as the topmost hazardous chemical in the world. Hence the reduction of nitrophenols gains the most attention. Generally, the nitrophenol reduction reaction is thermodynamically favourable (E0 = − 0.76 V) at optimized conditions, whereas the NaBH4 acts as a reducing agent (E0 = − 1.33 V)27,28. However, the reduction rate is prolonged without the catalyst due to the kinetic barrier between the reducing agent and reactant. Hence, the catalytic reduction nitrophenols are a good way to convert to non-toxic aminophenol (AP) with the presence of NaBH4 as a reducing agent. The reduction reaction was easily monitored with a UV–Vis spectrometer. It is known that NaBH4 alone cannot reduce the nitrophenols into aminophenol’s. As shown in Fig. SI. 2 the fresh nitrophenols (NP, DNP and TNP) absorption peak appeared at 300–370 nm respectively. When the addition of reducing agent, the peak was shifted to 402–450 nm28. In addition, the solution color was turned light yellow to deep yellow, due to the formation of corresponding nitrophenolate ions in basic solution. However, no reduction was achieved over 2 h, indicating that nitrophenolate ions were very stable with NaBH4. Furthermore, the catalytic activity of the NiO was explored with three nitrophenols and shown in Fig. SI. 3. The pure NiO exhibits a poor catalytic reduction of nitro compounds. In contrast, the PdO–NiO catalyst show excellent activity on the nitrophenols, as shown in Fig. 8.Figure 8Catalytic activity and kinetic rate of PdO–NiO nanocomposite on reduction of nitrophenols with NaBH4 solution (a,b) NP, (c,d) DNP and (e,f) TNP.Full size imageIt can be seen that the nitrophenolate peak absorbance at 400 nm gradually decreases with reaction time, which confirmed that PdO–NiO promotes the electron and hydrogen transfer between the reactant. Due to the higher active sites of the PdO–NiO. The present PdO–NiO catalyst completes the reduction reaction of NP, DNP and TNP within 10, 13, 25 min respectively. In addition, the aminophenol absorption peak appeared around 300 nm for all three nitrophenols. On the other hand, the deep yellow solution turned colorless, indicating the formation of aminophenol. The rate constant κapp for each nitrophenol was calculated from the plot of ln (At/Ao) Vs. time. The proposed PdO–NiO catalyst exhibits excellent rate constant 0.1667, 0.0997, 0.0686 min−1 for NP, DNP and TNT, respectively, which is the higher rate constant than the previously reported catalyst (Table SI. 2). Generally, the reduction mechanism of nitrophenols to aminophenols follows many intermediate steps from nitro to nitroso and then to hydroxylamine and to final aminophenol. For these reaction required both electron transfer and active hydrogen atoms. Here, the BH4− ions produce the active hydrogen atoms on the surface of the catalyst and subsequently, the PdO–NiO catalyst enhances the electron transfer. As a result, the reduction of NP could be efficiently accelerated by the PdO–NiO catalyst. Furthermore, comparison of the catalytic reduction performance of nitrophenols with varies catalyst are shown in Table SI. 2. The reduction mechanism of the nitrophenols with PdO–NiO catalyst with NaBH4 as shown in Fig. SI. 6.Furthermore, the catalytic activity of PdO–NiO composite was explored by the reduction of organic azo dye compounds such as Methylene blue (MB), Rhodamine B (RhB) and Methyl orange (MO) with the addition of NaBH4 in the presence of PdO–NiO catalyst29. The reduction of each dye was monitored at different absorption peaks, as shown in Fig. 9. The intensity of each dye at respective wavelengths linearly reduced with time in the presence of the PdO–NiO. In addition, the reduction rate κapp for each dye was calculated from the plot of ln (At/Ao) Vs. time. PdO–NiO catalyst exhibited excellent reduction rate as 0.099, 0.0416 and 0.0896 min−1 for MB, RhB and MO, respectively, superior catalytic activity than previous reports. In addition, the azo dye solution turned into colourless, indicating the complete reduction occurs in the presence of PdO–NiO. The pure NiO catalyst exhibits poor reduction activity towards azo dyes with the presence of NaBH4 (Fig. SI. 4). Furthermore, comparison of the catalytic reduction performance azo dyes with varies catalyst are shown in Table. SI. 3. Additionally, the reduction of the mixture of nitrophenols (NP, DNP and TNP) and azo dyes (MB, RhB and MO) was tested with PdO–NiO nanocomposite and obtained results are shown in Fig. 10. Initially, the mixture of azo compounds is formed dark solution then rapidly turned into a colourless and became a transparent solution with the addition of PdO–NiO in the presence of NaBH430. The complete azo compounds reduction was achieved within 8 min. Hence, the proposed PdO–NiO is a promising catalyst for wastewater treatment. In addition, the effect of the catalyst loading on the reduction of mixture of azo compounds were studied with different loading amount of PdO–NiO catalyst (3–10 mg) and shown in the Table. SI. 4. Which show that the reduction rate increased with loading amount of the catalyst. In addition, the catalytic reduction performance of toxic azo compounds by various catalysts are shown in Table. 2. PdO–NiO catalyst exhibit superior reduction performance than the previously reported catalyst.Figure 9Catalytic activity and kinetic rate of PdO–NiO nanocomposite on reduction of azo dyes with NaBH4 solution (a,b) MB, (c,d) RhB and (e,f) MO.Full size imageFigure 10Reaction progress of an azo compounds mixture (4-NP, 2,4-DNP, 2,4,6-TNP, MB, RhB, MO) with PdO–NiO and NaBH4. Conditions: Dye: 100 ppm, 25 ml, Nitrophenol: 0.12 mM, 25 ml, NaBH4: 0.1 M, 5 ml and catalyst: 3 mg.Full size imageTable 2 Comparison of catalytic reduction performance of toxic azo compounds by various catalysts.Full size tableFurthermore, the reduction mechanism of azo dyes over PdO–NiO catalyst with reducing agent shown in Fig. SI. 7.After the complete reduction reaction, the catalyst property was analyzed to understand the stability of PdO–NiO. In Fig. 11a, FTIR spectra showed no noticeable changes before and after catalytic reduction of mixture reduction. Additionally, the SEM image (Fig. 11b) also showed no changes in the morphology of the PdO–NiO. In addition, HR-TEM image (Fig. 11c) was also analyzed to study the change in the particles size after catalytic reduction, show no considerable change in the particle size. Which proved that PdO–NiO is highly stable in the reduction conditions.Figure 11(a) FTIR spectra of PdO–NiO nanocomposite before and after reducing the mixture of azo compounds (b) FE-SEM image (c) HR-TEM image of PdO–NiO after reducing the mixture of azo compounds.Full size image More

  • in

    Successful microbial colonization of space in a more dispersed manner

    Simulating competition for space using the “BacGo” modelTo investigate how spatial positioning of populations affects the outcome of microbial competition, we simulated two populations competing for space with a limiting size by building an individual-based model (named “BacGo”). The model was implemented in discrete grid boxes of a 20 × 20 array. As shown in Fig. 1a, our simulations were based on three basic assumptions. First, the two competing populations possess the same inherent growth rate and equal initial cell numbers, thus the only differences between them are their manners of colonizing free space. Second, the newly born daughter cell is located around its mother cell but with a random direction of spatial positioning [34], resulted in a microcolony with different spatial patterning. Lastly, if the selected box has been occupied, the newborn cell will compete for the box against the original occupants of the box and possesses a probability of 50% to survive [37].We first explored the outcome of spatial competition, which started by randomly distributing two populations on the grids with the same initial cell numbers of 10 for each. Based on our basic assumptions and the predictions of competitive exclusion theory [38], we hypothesized that only one population could win the competition and finally occupy all grids. As shown in 20,000 independent simulations with random initial distributions, we discovered that at the end of each simulation, only one population survived (Video S1 and Video S2). The Chi-square test showed no significant difference (P = 0.211) between the simulated winning times (10,177 of 20,000 simulations) and the random winning times (10,051 of 20,000 simulations) of the focus population. This result conformed with our initial assumption that cells possess a probability of 50% to survive in competing with original occupants. When we replicated simulations initiated with the same cell distribution, we found that the winning probabilities for each population changed in line with the initial distributions (Fig. S1). However, the winning probabilities never reached 100% no matter how the initial distribution changes. Together, these results suggested that unknown random factors may affect the final outcome of the competition.Next, we analyzed the dynamics of microbial colonization during our simulations. As summarized in Fig. 1b, we divided the competition process into two stages, the “occupation stage” and the “exclusion stage” (see Methods). To statistically characterize the competitive outcome at t3, we defined the winning asymmetry index, WinR, and the abundance asymmetry index, AbunR (see Methods). As shown in Fig. S2a, we found a strong positive correlation (R2 = 0.740, P  More

  • in

    Photoheterotrophy by aerobic anoxygenic bacteria modulates carbon fluxes in a freshwater lake

    SamplingWater was sampled from Cep lake in Czechia, at a regular sampling site of 10 m depth (48.944 °N, 14.877 °E). The lake originates from sand mining in the 1970–80s. It is a permanent meso-oligotrophic (chlorophyll-a concentrations ranged from 1.4–16.4 µg L−1) seepage reservoir filled with groundwater penetrating from the nearby river Lužnice. The lake area is about 1.16 km2, with the maximum depth about 11–12 m. These characteristics are representative for most of temperate and boreal lakes [16].Samples were collected every four weeks from April till October in 2018, and from April till November in 2019. Ten liters of water were collected from 0.5 m depth using a Ruttner Water Sampler (model 11.003KC Denmark AS). Temperature and oxygen profiles were taken with an EXO1 multi-parameter probe (YSI Inc., Yellow Springs, USA). Water was transported to the laboratory within 2 h from the collection in a closed container made from high-density polyethylene, rinsed three times with the sampled water and stored in a cooled box.NutrientsSamples were filtered through glass fiber filters with 0.4 µm nominal porosity (GF-5, Macherey-Nagel, Düren, Germany). Concentrations of soluble reactive phosphorus (SRP) were determined spectrophotometrically [17, 18]. Concentrations of nitrate and ammonium were measured according to Procházková [19] and Kopáčkek and Procházková [20]. Dissolved organic carbon (DOC) and dissolved nitrogen (DN) were measured with a TOC 5000 A analyzer (Shimadzu, Kyoto, Japan).PigmentsSeston from 1.43 to 3.65 L of water was collected onto GF-5 glass fiber filters (diameter 47 mm, Macherey-Nagel). The filters were dried of excess water by gently pressing in a paper towel, and flush frozen in liquid nitrogen. Pigments were extracted and analyzed by HPLC as described in Piwosz et al. [21].Net primary production (NPP) and community respiration (oxygen measurements)Oxygen concentration was measured with the Winkler method [22]. It was chosen because it allows O2 concentration to be measured directly in the water without the need to consider carbonic equilibrium, which is the case when changes in CO2 concentration is measured [23]. Samples were unfiltered to avoid the removal of particle-associated bacteria and also of free-living AAP bacteria, which tend to be larger than average freshwater bacteria [12]. Glass stoppered Winkler type oxygen bottles (115 mL nominal capacity, VTR glass, Prague, Czechia) were filled with the sampled water directly from the sampler via a rubber tube. Each bottle was first rinsed three times and then filled without the formation of air bubbles. Water was allowed to overflow the neck of the bottle for about 1 min, and the bottle was closed with a glass stopper to avoid air bubbles. The closed bottles were kept in the dark in a cooled box. On the shore, three bottles were selected as T0, and 1.2 mL of manganese (II) chloride solution (concentration 3 mol L−1) was addded, followed by the addition of 1.2 mL of a mixture containing 4 mol L−1 of sodium iodide solution and 8 mol L−1 of sodium hydroxide solution. These samples were processed in the laboratory within 3 h. The remaining bottles were incubated for 24 h at in situ temperature in the IR-box prepared from the MAKROLON IR polycarbonate sheet (4 mm thickness, Professional Plastics, Inc. Fullerton, CA, USA). These panels have a maximum transmittance of 90% in the infrared region 850–2000 nm, 50% at 780 nm, and 0% 10 °C according to our measurements (180 days, Equation 3 in Supplementary File S1). Subsequently, the differences in the Cep Lake’s carbon budget for the surface layer (down to 0.5 m depth) were calculated by multiplying the integrated values by the volume of this layer (Equation 4 and 5 in Supplementary File S1).HCO3
    − incorporationTriplicated water samples (32 mL) were incubated for 3.2–5.2 h in the IR light and the dark at in situ temperature, as described for respiration. Total activity added to each bottle was measured from 1 mL aliquot of the incubated sample that was transferred to a scintillation vial containing 20 μl of 5 mol L−1 NaOH (to prevent a loss of 14C-bicarbonate). Thirty mL of sample was filtered through 2.5 μm nitrate cellulose filters (Pragopor, Prague, Czechia, diameter 25 mm). Five mL of the filtered water was collected and subsequently filtered through a 0.17 μm nitrate cellulose filter. The resulting cell-free filtrate, which contained 14C-DOC was collected. The filtration was done at a low vacuum (0.02 MPa) to avoid cell breakage. The total CO2 assimilation rate was calculated as the sum of all these fractions.The filters were kept in an HCl-saturated atmosphere for 24 h at room temperature in a fume hood. They were placed in scintillation vials and dissolved in 1 mL of ethyl acetate (Penta, Prague, Czechia). Then, 5 mL of Ultima Golt LLT scintillation cocktail (PerkinElmer, Waltham, MA, USA) was added. Five mL of cell-free filtrates were acidified by adding 100 μL 5 mol L−1 HCl to volatilize non-incorporated H14CO3 and incubated 24 h at room temperature in a fume hood. Then, 10 mL of the scintillation cocktail was added. Finally, 5 mL of the scintillation cocktail was added to the total activity samples. Subsequently, the samples were gently mixed and left in the dark for 48 h. The radioactivity in the samples was measured using a Tri-Carb 2810 TR scintillation counter (PerkinElmer).To estimate carbon fluxes (μmol C L–1 h–1), a fraction of the added H14CO3 incorporated or released was corrected for the incubation time and multiplied by the concentration of total dissolved inorganic carbon (DIC). The DIC concentration was calculated based on temperature, pH, and alkalinity measurements (Inolab pH 720, WTW Xylem Inc. Rye Brook, NY, US) determined by Gran titration.Assimilation of organic monomersThe difference between microbial activity in the IR light and dark was also estimated based on assimilation rates of radiolabeled glucose, pyruvate, leucine and thymidine (American Radiolabeled Chemicals, St. Louis, MO, USA). Tritiated glucose (specific activity (SA): 2220 GBq mmol−1), leucine (SA: 4440 GBq mmol−1) and thymidine (SA: 2275.5 GBq mmol−1) were added to 5 mL samples to a final concentration of 5 nmol L−1. 14C-pyruvate (SA: 2.035 GBq mmol−1) was added to a final concentration of 10 nmol L−1. Trichloroacetic acid (TCA) was added to the killed controls to a final concentration of 1%. Samples were incubated for 1 h in the dark and IR light as described for respiration. The incubations were terminated as the killed controls and kept at 4 °C in the dark until processed within 65% except for the samples from 9th May and 29th Aug 2018 (10% each), 1st Aug 2018 (21%), 25th Nov 2018 (1%), and 14th Aug 2019 (22%), Supplementary Fig. S1A). Thus, we decided to concatenate the fastaq files and analyze both fractions together as the total community. This also facilitated statistical analysis, as the activity rates were measured for the whole community without fractionations.Reads quality was evaluated using FastQC v0.11.7 (Babraham Bioinformatics, Cambridge, UK). After primer sequences trimming using Cutadapt [29] (v1.16), the number of reads per sample ranged from 49,354 to 188,942. Subsequent analyses were done in the R/Bioconductor environment using the dada2 package (version 1.14.1) [30]. Forward and reverse reads were truncated to 225 bp and low quality sequences were filtered out with the filterAndTrim function (truncLen = c(225, 225), maxN = 0, maxEE = c(2, 2), truncQ = 2), which reduced the number of reads per sample to range from 30,190 to 143,552. After merging and chimera removal using the removeBimeraDenovo function, 4,893 amplicon sequence variants (ASV) were obtained. Rare ASVs (not seen >3 times in at least 20% of the samples) were removed, which reduced the number of ASVs to 658, and the number of reads to 14,613–69,046 per sample. Taxonomic assignment was done using SILVA 138.1 database [31, 32] released on August 27, 2020. ASVs identified as Chloroplast or Cyanobacteria were excluded from the analyses, giving the final number of 546 ASVs and from 10,819 to 54,799 reads per sample. The bacterial community composition graphs were done using phyloseq [33] and ggplot2 [34] packages.AAP community compositionThe composition of AAP community was analyzed by amplicon sequencing of pufM gene encoding the M subunit of bacterial type-2 reaction centers. This gene is routinely used for diversity studies of AAP bacteria [35].PufM gene amplicons (approx. 245 bp) were prepared using pufM_UniF (5′-GGN AAY YTN TWY TAY AAY CCN TTY CA-3′) and pufM_WAW (5′-AYN GCR AAC CAC CAN GCC CA-3′) primers [36]. PCR was performed in triplicate 20 μL reactions using Phusion™ High-Fidelity DNA Polymerase (Thermo Scientific, USA) with the following reaction conditions: 98 °C for 3 min, 27 cycles at 98 °C for 10 s, 58 °C for 30 s, 72 °C for 30 s, and a final extension at 72 °C for 5 min. The triplicate product reactions for each sample were pooled and gel purified using the kit Wizard SV Gel and PCR Clean-Up System (Promega, USA). The sequencing was performed on the Illumina MiSeq platform (2 × 150 bp) at Macrogen, South Korea.The fastq files were concatenated as described for bacteria communities. The Bray-Curtis similarity between two fractions for each sampling day was >70%, except for the samples from 1st Aug 2018 (47%) and 14th Aug 2019 (18%, Supplementary Fig. S1B).The samples were analyzed as described for bacterial communities. The number of reads per sample ranged from 192,360 to 239,418 after the cutadapt trimming. Forward and reverse reads were truncated to 130 bp, and the number of reads per sample after the quality filtering and denoising ranged from 189,432 to 235,311. Merging the forward and reverse reads with mergePairs function created 12,692 ASVs and reduced the number of reads to 183,136–221,281 per sample. The chimera removal lowered the number of ASVs to 1816, and the number of reads to 159,451–208,679. Rare ASVs (not seen >3 times in at least 20% of the samples) were removed, which resulted in the final 566 ASVs, and a number of reads ranging from 155,915 to 203,021 per sample. A manually curated taxonomic database was used for taxonomic assignment following the naïve Bayesian classifier method [37]. It contained 1580 unique pufM sequences, downloaded from the Fungene repository on May 16, 2019 (http://fungene.cme.msu.edu [38]), from metagenomes from the Římov Reservoir [39, 40] and from the Genome Taxonomy database accessed on September 16, 2020 [41].Statistical analysisLinear mixed-effects models were calculated in R (version 3.6.2) using lme function from the nlme package (version 3.1.143) on untransformed activity data and log10 transformed environmental variables [42]. Models’ parameters were estimated using maximum likelihood method and their significance was tested with ANOVA. Relationships between the activity measures, the environmental variables and the composition of AAP communities were investigated with distance-based linear models (DistLM) [43, 44] in Primer (version 7.0.13) with PERMANOVA + 1 add on (e-Primer, Plymouth, UK) [45]. The sequence reads were transformed with the varianceStabilizingTransformation function of the DESeq2 package [46] (version 1.14.1, blind = FALSE, fitType = “mean”).Data accessibilityThe sequences of 16S and pufM amplicons that support the findings of this study have been deposited in the EMBL database as the BioProject with the accession number PRJEB41596, together with most of the environmental metadata. The scripts and the remaining data supporting the results are included in the Supplementary Material. More

  • in

    Direct and indirect effects of roads on space use by jaguars in Brazil

    1.Hughes, A. C. et al. Horizon scan of the belt and road initiative. Trends Ecol. Evol. 35(7), 583–593. https://doi.org/10.1016/j.tree.2020.02.005 (2020).Article 
    PubMed 

    Google Scholar 
    2.Meijer, J. R., Huijbregts, M. A. J., Schotten, K. C. G. J. & Schipper, A. M. Global patterns of current and future road infrastructure. Environ. Res. Lett. 13(6), 064006. https://doi.org/10.1088/1748-9326/aabd42 (2018).ADS 
    Article 

    Google Scholar 
    3.DNIT. Sistema de Gerencia de Pavimentos. Relatório dos levantamentos funcionais das rodovias federais. Departamento Nacional de Infraenstrutura de Transporte (2013). www.dnit.gov.br.4.Teixeira, F. et al. The need to improve and integrate science and environmental licensing to mitigate wildlife mortality on roads in Brazil. Trop. Conserv. Sci. 9, 24–42. https://doi.org/10.1177/194008291600900104 (2016).Article 

    Google Scholar 
    5.Barber, C. P., Cochrane, M. A., Souza, C. M. & Laurance, W. F. Roads, deforestation, and the mitigating effect of protected areas in the Amazon. Biol. Conserv. 177, 203–209. https://doi.org/10.1016/j.biocon.2014.07.004 (2014).Article 

    Google Scholar 
    6.Reid, J. & Souza, W. C. Infrastructure and conservation policy in Brazil. Conserv. Biol. 19, 740–746. https://doi.org/10.1111/j.1523-1739.2005.00699.x (2005).Article 

    Google Scholar 
    7.Bowman, J., Ray, C. R., Magoun, A. J. & Johnson, D. F. N. Roads, logging, and the large-mammal community of an eastern Canadian boreal forest. Can. J. Zool. 88(5), 454–467. https://doi.org/10.1139/z10-019 (2010).Article 

    Google Scholar 
    8.Fahrig, L. & Rytwinski, T. Effects of roads on animal abundance: An empirical review and synthesis. Ecol. Soc. 14(1), 21 (2009).Article 

    Google Scholar 
    9.Laurance, W. F., Goosem, M. & Laurance, S. G. W. Impacts of roads and linear clearings on tropical forests. Trends Ecol. Evol. 24, 659–669. https://doi.org/10.1016/j.tree.2009.06.009 (2009).Article 
    PubMed 

    Google Scholar 
    10.Ruiz-Capillas, P., Mata, C. & Malo, J. E. Road verges are refuges for small mammal populations in extensively managed Mediterranean landscapes. Biol. Conserv. 158, 223–229. https://doi.org/10.1016/j.biocon.2012.09.025 (2013).Article 

    Google Scholar 
    11.Grilo, C. et al. Individual spatial responses towards roads: Implications for mortality risk. PLoS ONE 7(9), e43811. https://doi.org/10.1371/journal.pone.0043811 (2012).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    12.Jacobson, S. L., Bliss-Ketchum, L. L., de Rivera, C. E. & Smith, W. P. A behavior-based framework for assessing barrier effects to wildlife from vehicle traffic volume. Ecosphere 7, e01345. https://doi.org/10.1002/ecs2.1345 (2016).Article 

    Google Scholar 
    13.Li, T. et al. Fragmentation of China’s landscape by roads and urban areas. Landsc. Ecol. 25, 839–853. https://doi.org/10.1007/s10980-010-9461-6 (2010).Article 

    Google Scholar 
    14.Walker, R. et al. Modeling spatial decisions with graph theory: Logging roads and forest fragmentation in the Brazilian Amazon. Ecol. Appl. 23, 239–254. https://doi.org/10.1890/11-1800.1 (2013).Article 
    PubMed 

    Google Scholar 
    15.Fahrig, L. Effects of habitat fragmentation on biodiversity. Annu. Rev. Ecol. Evol. Syst. 34, 487–515. https://doi.org/10.1146/annurev.ecolsys.34.011802.132419 (2003).Article 

    Google Scholar 
    16.Püttker, T. et al. Indirect effects of habitat loss via habitat fragmentation: A cross-taxa analysis of forest-dependent species. Biol. Conserv. 241, 108368. https://doi.org/10.1016/j.biocon.2019.108368 (2020).Article 

    Google Scholar 
    17.Signorelli, L., Bastos, R. P., De Marco, P. & With, K. A. Landscape context affects site occupancy of pond-breeding anurans across a disturbance gradient in the Brazilian Cerrado. Landsc. Ecol. 31, 1997. https://doi.org/10.1007/s10980-016-0376-8 (2016).Article 

    Google Scholar 
    18.Zimmermann, B., Nelson, L., Wabakken, P., Sand, H. & Liberg, O. Behavioral responses of wolves to roads: Scale-dependent ambivalence. Behav. Ecol. 25, 1353–1364. https://doi.org/10.1093/beheco/aru134 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    19.Jaeger, J. A. G. et al. Predicting when animal populations are at risk from roads: An interactive model of road avoidance behaviour. Ecol. Model. 185, 329–348. https://doi.org/10.1016/j.ecolmodel.2004.12.015 (2005).Article 

    Google Scholar 
    20.Jaeger, J. A. G., Fahrig, L., Ewald, K. C. Does the configuration of road networks influence the degree to which roads affect wildlife populations? In Proceedings of the 2005 International Conference on Ecology and Transportation (ICOET) (eds Irwin, C. L., Garrett, P. & McDermott, K. P.) 151–163 (Center for Transportation and the Environment, North Carolina State University, Raleigh, NC, 2006).21.Rytwinski, T. & Fahrig, L. Reproductive rates and body size predict road impacts on mammal abundance. Ecol. Appl. 21(2), 589–600. https://doi.org/10.1890/10-0968.1 (2011).Article 
    PubMed 

    Google Scholar 
    22.Rytwinski, T. & Fahrig, L. Do species life history traits explain population responses to roads? A meta-analysis. Biol. Conserv. 147(1), 87–98. https://doi.org/10.1016/j.biocon.2011.11.023 (2012).Article 

    Google Scholar 
    23.Nowell, K. & Jackson, P. Wilds Cats: Status Survey e Conservation Action Plan (IUCN, 1996).
    Google Scholar 
    24.de La Torre, J., Zarza, H., Ceballos, G. & Medellin, R. The jaguars’ spots are darker than they appear: Assessing the global conservation status of the jaguar Panthera onca. Oryx 51, 1–16. https://doi.org/10.1017/S0030605316001046 (2017).Article 

    Google Scholar 
    25.Morrison, J. C., Sechrest, W., Dinerstein, E., Wilcove, D. S. & Lamoreux, J. F. Persistence of large mammal faunas as indicators of global human impacts. J. Mammal. 88(6), 1363–1380. https://doi.org/10.1644/06-MAMM-A-124R2.1 (2007).Article 

    Google Scholar 
    26.Alvarenga, G. C. et al. Multi-scape path-level analysis of jaguar habitat use in the Pantanal ecosystem. Biol. Conserv. 253, 108900. https://doi.org/10.1016/j.biocon.2020.108900 (2021).Article 

    Google Scholar 
    27.Espinosa, S., Celis, G. & Branch, L. C. When roads appear jaguars decline: Increased access to an Amazonian wilderness area reduces potential for jaguar conservation. PLoS ONE 13(1), e0189740. https://doi.org/10.1371/journal.pone.0189740 (2018).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    28.Pallares, E., Manterolla, C., Conde, D. A. & Colchero, F. Case Study: roads and jaguars in the Mayan Forest. In Handbook of Road Ecology (eds Van der Ree, R. et al.) 313–136 (Wiley, 2015).
    Google Scholar 
    29.Zeilhofer, P., Cezar, A., Tôrres, N. M., Jácomo, A. T. A. & Silveira, L. Jaguar Panthera onca habitat modeling in landscapes facing high land-use transformation pressure—Findings from Mato Grosso, Brazil. Biotropica 46(1), 98–105. https://doi.org/10.1111/btp.12074 (2014).Article 

    Google Scholar 
    30.Colchero, F., Conde, D. A., Manterola, C., Rivera, A. & Ceballos, G. Jaguars on the move: Modeling movement to mitigate fragmentation from road expansion in the Mayan Forest. Anim. Conserv. 4, 158–166. https://doi.org/10.1111/j.1469-1795.2010.00406.x (2011).Article 

    Google Scholar 
    31.Conde, D. A. et al. Sex matters: Modeling male and female habitat differences for jaguar conservation. Biol. Conserv. 143(9), 1980–1988. https://doi.org/10.1016/j.biocon.2010.04.049 (2010).Article 

    Google Scholar 
    32.De Angelo, C., Paviolo, A., Wiegand, T., Kanagaraj, R. & Di Bitetti, M. S. Understanding species persistence for defining conservation actions: A management landscape for jaguars in the Atlantic Forest. Biol. Conserv. 159, 422–433. https://doi.org/10.1016/j.biocon.2012.12.021 (2013).Article 

    Google Scholar 
    33.Grace, J. B. Structural Equation Modeling and Natural Systems (Cambridge University Press, 2006).Book 

    Google Scholar 
    34.Shipley, B. Confirmatory path analysis in a generalized multilevel context. Ecology 90, 363–368. https://doi.org/10.1890/08-1034.1 (2009).Article 

    Google Scholar 
    35.Morato, R. G. et al. Jaguar movement database: A GPS-based movement dataset of an apex predator in the Neotropics. Ecology 99, 1691–1691. https://doi.org/10.1002/ecy.2379 (2018).Article 
    PubMed 

    Google Scholar 
    36.Geofabrik. OpenStreetMap-Shapefiles (2019). http://download.geofabrik.de. Accessed 15 Aug 2019.37.Projeto MapBiomas – Coleção 2 da Série Anual de Mapas de Cobertura e Uso de Solo do Brasil, acessado em 12/04/2019 através do link. http://mapbiomas.org/pages/database/mapbiomas_collection_download.38.Morato, R. G. et al. Resource selection in an apex predator and variation in response to local landscape characteristics. Biol. Conserv. 228, 233–240. https://doi.org/10.1016/j.biocon.2018.10.022 (2018).Article 

    Google Scholar 
    39.ESRI. Environmental Systems Research Institute. ArcGIS. Geographic Information System for Desktop, version 10.3.1 (2015).40.RStudio Team. RStudio: Integrated Development for R (RStudio Inc, 2016).
    Google Scholar 
    41.Lefcheck, J. S. piecewiseSEM: Piecewise structural equation modelling in R for ecology, evolution, and systematics. Methods Ecol. Evol. 7, 573–579. https://doi.org/10.1111/2041-210X.12512 (2016).Article 

    Google Scholar 
    42.Grace, J. B., Adler, P. B., Harpole, W. S., Borer, E. T. & Seabloom, E. W. Causal networks clarify productivity–richness interrelations, bivariate plots do not. Funct. Ecol. 28, 787–798. https://doi.org/10.1111/1365-2435.12269 (2014).Article 

    Google Scholar 
    43.Bollen, K. A. & Pearl, J. Eight myths about causality and structural equation models. In Handbook of Causal Analysis for Social Research (ed. Morgan, S. L.) 301–328 (Springer, 2013).Chapter 

    Google Scholar 
    44.Fan, Y. et al. Applications of structural equation modeling (SEM) in ecological studies: An updated review. Ecol. Process 5, 19. https://doi.org/10.1186/s13717-016-0063-3 (2016).Article 

    Google Scholar 
    45.Cressie, N. A. C. Statistics for Spatial Data. Wiley Series in Probability and Mathematical Statistics (Wiley, 1993).
    Google Scholar 
    46.Haining, R. Spatial Data Analysis: Theory and Practice (Cambridge University Press, 2003).Book 

    Google Scholar 
    47.Amrhein, V., Greenland, S. & McShane, B. Retire statistical significance. Nature 567(7748), 305–307. https://doi.org/10.1038/d41586-019-00857-9 (2019).ADS 
    CAS 
    Article 

    Google Scholar 
    48.Benítez-López, A., Alkemade, R. & Verweij, P. A. The impacts of roads and other infrastructure on mammal and bird populations: A meta-analysis. Biol. Conserv. 143, 1307–1316. https://doi.org/10.1016/j.biocon.2010.02.009 (2010).Article 

    Google Scholar 
    49.Torres, A., Jaeger, J. A. & Alonso, J. C. Assessing large-scale wildlife responses to human infrastructure development. Proc. Natl. Acad. Sci. USA 113(30), 8472–8477. https://doi.org/10.1073/pnas.1522488113 (2016).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    50.Morato, R. G., Ferraz, K. M. P. M. B., Paula, R. C. & Campos, C. B. Identification of priority conservation areas and potential corridors for jaguars in the Caatinga biome, Brazil. PLoS ONE 9(4), e92950. https://doi.org/10.1371/journal.pone.0092950 (2014).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    51.Rodríguez-Soto, C. et al. Predicting potential distribution of the jaguar (Panthera onca) in Mexico: Identification of priority areas for conservation. Divers. Distrib. 17, 350–361. https://doi.org/10.1111/j.1472-4642.2010.00740.x (2011).Article 

    Google Scholar 
    52.Ripple, W. J. et al. Status and ecological effects of the world’s largest carnivores. Science 343(6167), 1241484. https://doi.org/10.1126/science.1241484 (2014).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    53.Rabinowitz, A. & Zeller, K. A. A range-wide model of landscape connectivity and conservation for the jaguar, Panthera onca. Biol. Conserv. 143(4), 939–945. https://doi.org/10.1016/j.biocon.2010.01.002 (2010).Article 

    Google Scholar 
    54.Jaeger, J. Landschaftszerschneidung. Eine transdisziplinäre Studie gemäß dem Konzept der Umweltgefährdung (Landscape Fragmentation. A Transdisciplinary Study According to the Concept of Environmental Threat) (Eugen Ulmer, 2002).
    Google Scholar 
    55.Torres, A., Jaeger, J. A. G. & Alonso, J. C. Multi-scale mismatches between urban sprawl and landscape fragmentation create windows of opportunity for conservation development. Landsc. Ecol. 31, 2291–2305. https://doi.org/10.1007/s10980-016-0400-z (2016).Article 

    Google Scholar 
    56.Morato, R. G. et al. space use and movement of a neotropical top predator: The endangered jaguar. PLoS ONE 11(12), e0168176. https://doi.org/10.1371/journal.pone.0168176 (2016).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    57.Sollmann, R. et al. Improving density estimates for elusive carnivores: Accounting for sex-specific detection and movement using spatial capture-recapture models for jaguars in central Brazil. Biol. Conserv. 144, 1017–1024. https://doi.org/10.1016/j.biocon.2010.12.011 (2011).Article 

    Google Scholar 
    58.Marchini, S. & Macdonald, D. W. Mind over matter: Perceptions behind the impact of jaguars on human livelihoods. Biol. Conserv. 224, 230–237. https://doi.org/10.1016/j.biocon.2018.06.001 (2018).Article 

    Google Scholar 
    59.González-Gallina, A. et al. Home range of a male jaguar spatially associated with the landfill of the city of Playa del Carmen, Mexico. Mammalia 82(1), 54–61. https://doi.org/10.1515/mammalia-2016-0065 (2017).Article 

    Google Scholar 
    60.Associação Onçafari. Enteda o caso da onça-pintada de Juiz de Fora (2019). https://oncafari.org/2019/06/03/entenda-o-caso-da-onca-pintada-de-juiz-de-fora/. Accessed 09 Oct 2020.61.Dickson, B. G., Jenness, J. S. & Beier, P. Influence of vegetation, topography, and roads on cougar movement in southern California. J. Wildl. Manag. 69, 264–276. https://doi.org/10.2193/0022-541X(2005)069%3c0264:IOVTAR%3e2.0.CO;2 (2005).Article 

    Google Scholar 
    62.Thatte, P., Joshi, A., Vaidyanathan, S., Landguth, E. & Ramakrishnan, U. Maintaining tiger connectivity and minimizing extinction into the next century: Insights from landscape genetics and spatially-explicit simulations. Biol. Conserv. 218, 181–191. https://doi.org/10.1016/j.biocon.2017.12.022 (2018).Article 

    Google Scholar 
    63.Thompson, J. J. et al. Environmental and anthropogenic factors synergistically affect space use of jaguars. Curr. Biol. 31(15), 3457–3466. https://doi.org/10.1016/j.cub.2021.06.029 (2021).CAS 
    Article 
    PubMed 

    Google Scholar 
    64.Dickson, B. G. & Beier, P. Home-range and habitat selection by adult cougars in southern California. J. Wildl. Manag. 66, 1235–1245. https://doi.org/10.2307/3802956 (2002).Article 

    Google Scholar 
    65.Poessel, S. A. et al. Roads influence movement and home ranges of a fragmentation-sensitive carnivore, the bobcat, in an urban landscape. Biol. Conserv. 180, 224–232. https://doi.org/10.1016/j.biocon.2014.10.010 (2014).Article 

    Google Scholar 
    66.Johnson, D. H. The comparison of usage and availability measurements for evaluation of resource preference. Ecology 61, 65–71. https://doi.org/10.2307/1937156 (1980).Article 

    Google Scholar 
    67.Silva, L. G., Cherem, J., Kasper, C., Trigo, T. & Eizirik, E. Mapping wild cat roadkills in southern Brazil: An assessment of baseline data for species conservation. Cat News (Bougy) 61, 04–07. https://doi.org/10.13140/RG.2.2.17640.88327 (2014).Article 

    Google Scholar 
    68.Srbek-Araujo, A. C., Mendes, S. L. & Chiarello, A. G. Jaguar (Panthera onca Linnaeus, 1758) roadkill in Brazilian Atlantic Forest and implications for species conservation. Braz. J. Biol. 75, 581–586. https://doi.org/10.1590/1519-6984.17613 (2015).CAS 
    Article 
    PubMed 

    Google Scholar 
    69.Kerley, L. L. et al. Effects of roads and human disturbance on Amur tigers. Conserv. Biol. 16, 97–108. https://doi.org/10.1046/j.1523-1739.2002.99290.x (2002).Article 

    Google Scholar 
    70.Teixeira, F. Z., Rytwinski, T. & Fahrig, L. Inference in road ecology: What we know versus what we think we know. Biol. Lett. https://doi.org/10.1098/rsbl.2020.0140 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    71.Laurance, W. F. The anthropocene. Curr. Biol. 29, 942–995. https://doi.org/10.1016/j.cub.2019.07.055 (2019).CAS 
    Article 

    Google Scholar 
    72.Kanda, C. Z. et al. Spatiotemporal dynamics of conspecific movement explain a solitary carnivore’s space use. J. Zool. 308, 66–74. https://doi.org/10.1111/jzo.12655 (2019).Article 

    Google Scholar 
    73.Zarco-González, M. M., Monroy-Vilchis, O. & Alaníz, J. Spatial model of livestock predation by jaguar and puma in Mexico: Conservation planning. Biol. Conserv. 159, 80–87. https://doi.org/10.1016/j.biocon.2012.11.007 (2013).Article 

    Google Scholar 
    74.Bager, A., Borghi, C. E. & Secco, H. The influence of economics, politics, and environment on road ecology in South America. In Handbook of Road Ecology (eds Van der Ree, R. et al.) 407–413 (Wiley, 2015).
    Google Scholar 
    75.Kaszta, Z. et al. Simulating the impact of Belt and Road initiative and other major developments in Myanmar on an ambassador felid, the clouded leopard, Neofelis nebulosa. Landsc. Ecol. 35, 727–746. https://doi.org/10.1007/s10980-020-00976-z (2020).Article 

    Google Scholar 
    76.Cullen, L. Jr. et al. Implications of fine-grained habitat fragmentation and road mortality for jaguar conservation in the Atlantic Forest, Brazil. PLoS ONE 11(12), e0167372. https://doi.org/10.1371/journal.pone.0167372 (2016).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    77.Ceia-Hasse, A., Borda-de-Agua, L., Grilo, C. & Pereira, H. M. Global exposure of carnivores to roads. Glob. Ecol. Biogeogr. 26, 592–600. https://doi.org/10.1111/geb.12564 (2017).Article 

    Google Scholar 
    78.Grilo, C., Koroleva, E., Andrášik, R., Bíl, M. & González-Suárez, M. Roadkill risk and population vulnerability in European birds and mammals. Front. Ecol. Environ. 18(6), 323–328. https://doi.org/10.1002/fee.2216 (2020).Article 

    Google Scholar 
    79.Cerqueira, R. C. et al. Potential movement corridor and high road-kill likelihood do not spatially coincide for felids in Brazil: Implications for road mitigation. Environ. Manag. 67, 412–423 (2021).Article 

    Google Scholar 
    80.Clevenger, A. P. & Waltho, N. Performance indices to identify attributes of highway crossing structures facilitating movement of large mammals. Biol. Conserv. 121, 453–464. https://doi.org/10.1016/j.biocon.2004.04.025 (2005).Article 

    Google Scholar 
    81.Spanowicz, A. G., Teixeira, F. Z. & Jaeger, J. A. G. An adaptive plan for prioritizing road sections for fencing to reduce animal mortality. Conserv. Biol. 34(5), 1210–1220. https://doi.org/10.1111/cobi.13502 (2020).Article 
    PubMed 

    Google Scholar 
    82.Laurance, W. F. et al. A global strategy for road building. Nature 513, 229–232. https://doi.org/10.1038/nature13717 (2014).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    83.Carter, N., Killion, A., Easter, T., Brandt, J. & Ford, A. Road development in Asia: Assessing the range-wide risks for tigers. Sci. Adv. 6(18), eaaz9619. https://doi.org/10.1126/sciadv.aaz9619 (2020).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    84.Galetti, M. et al. Atlantic rainforest’s jaguars in decline. Science 342, 930. https://doi.org/10.1126/science.342.6161.930-a (2013).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    85.Paviolo, A. et al. A biodiversity hotspot losing its top predator: The challenge of jaguar conservation in the Atlantic Forest of South America. Sci. Rep. 6, 37147. https://doi.org/10.1038/srep37147 (2016).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    86.Freitas, S. R., Hawbaker, T. J. & Metzger, J. P. Effects of roads, topography, and land use on forest cover dynamics in the Brazilian Atlantic Forest. For. Ecol. Manag. 259, 410–417. https://doi.org/10.1016/j.foreco.2009.10.036 (2010).Article 

    Google Scholar  More

  • in

    Spatial distribution of conspecific genotypes within chimeras of the branching coral Stylophora pistillata

    1.Rinkevich, B. & Weissman, I. L. Chimeras in colonial inverebrates: A synergistic symbiosis or somatic- and cell-germ parasitism? Symbiosis 4, 117–134 (1987).
    Google Scholar 
    2.Buss, L. W. Somatic cell parasitism and the evolution of somatic tissue compatibility. Proc. Natl. Acad. Sci. U.S.A. 79, 5337–5341. https://doi.org/10.1073/pnas.79.17.5337 (1982).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    3.Foster, K. R., Fortunato, A., Strassmann, J. E. & Queller, D. C. The costs and benefits of being a chimera. Proc. R. Soc. Lond. B 269, 2357–2362. https://doi.org/10.1098/rspb.2002.2163 (2002).Article 

    Google Scholar 
    4.Money, N. P. Fungal get-together. Nature 405, 751. https://doi.org/10.1038/35015659 (2000).CAS 
    Article 
    PubMed 

    Google Scholar 
    5.Franks, T., Botta, R., Thomas, M. & Franks, J. Chimerism in grapevines: Implications for cultivar identity, ancestry and genetic improvement. Theor. Appl. Genet. 104, 192–199. https://doi.org/10.1007/s001220100683 (2002).CAS 
    Article 
    PubMed 

    Google Scholar 
    6.Casares, A. & Sylvain, F. F. Higher reproductive success for chimeras than solitary individuals in the kelp Lessonia spicata but no benefit for individual genotypes. Evol. Ecol. 30, 953–972. https://doi.org/10.1007/s10682-016-9849-0 (2016).Article 

    Google Scholar 
    7.Santelices, B., González, A. V., Beltrán, J. & Flores, V. Coalescing red algae exhibit noninvasive, reversible chimerism. J. Phycol. 53, 59–69. https://doi.org/10.1111/jpy.12476 (2017).CAS 
    Article 
    PubMed 

    Google Scholar 
    8.Gauthier, M. & Degnan, B. M. Partitioning of genetically distinct cell populations in chimeric juveniles of the sponge Amphimedon queenslandica. Dev. Comp. Immunol. 32, 1270–1280. https://doi.org/10.1016/j.dci.2008.04.002 (2008).CAS 
    Article 
    PubMed 

    Google Scholar 
    9.Fidler, A. E., Bacq-Labreuil, A., Rachmilovitz, E. & Rinkevich, B. Efficient dispersal and substrate acquisition traits in a marine invasive species via transient chimerism and colony mobility. PeerJ 2018, 1–23. https://doi.org/10.7717/peerj.5006 (2018).CAS 
    Article 

    Google Scholar 
    10.Rinkevich, B. & Weissman, I. Chimeras in colonial invertebrates: A synergistic symbiosis or somatic-and germ-cell parasitism. Symbiosis 4, 117–134 (1987).
    Google Scholar 
    11.Amar, K.-O., Chadwick, N. E. & Rinkevich, B. Coral kin aggregations exhibit mixed allogeneic reactions and enhanced fitness during early ontogeny. BMC Evol. Biol. 8, 126–126. https://doi.org/10.1186/1471-2148-8-126 (2008).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    12.Puill-Stephan, E., Willis, B., van Herwerden, L. & van Oppen, M. Chimerism in wild adult populations of the broadcast spawning coral Acropora millepora on the Great Barrier Reef. PLoS One 4, e7751. https://doi.org/10.1371/journal.pone.0007751 (2009).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    13.Hoeg, J. T. & Lutzen, J. Life cycle and reproduction in the Cirripedia, Rhizocephala. Oceanogr. Mar. Biol. Annu. Rev. 33, 427–485 (1995).
    Google Scholar 
    14.Gianasi, B. L., Hamel, J. F. & Mercier, A. Full allogeneic fusion of embryos in a holothuroid echinoderm. Proc. R. Soc. Lond. B 285, 1–7. https://doi.org/10.1098/rspb.2018.0339 (2018).CAS 
    Article 

    Google Scholar 
    15.Rinkevich, B. Human natural chimerism: An acquired character or a vestige of evolution?. Hum. Immunol. 62, 651–657. https://doi.org/10.1016/S0198-8859(01)00249-X (2001).CAS 
    Article 
    PubMed 

    Google Scholar 
    16.Gill, D. E., Chao, L., Perkins, S. L. & Wolf, J. B. Genetic mosaicism in plants and clonal animals. Annu. Rev. Ecol. Syst. 26, 423–444 (1995).Article 

    Google Scholar 
    17.Biesecker, L. G. & Spinner, N. B. A genomic view of mosaicism and human disease. Nat. Rev. Genet. 14, 307–320 (2013).CAS 
    Article 

    Google Scholar 
    18.Devlin-Durante, M. K., Miller, M. W., Precht, W. F. & Baums, I. B. How old are you? Genet age estimates in a clonal animal. Mol. Ecol. 25, 5628–5646. https://doi.org/10.1111/mec.13865 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    19.Dubé, C. E., Planes, S., Zhou, Y., Berteaux-Lecellier, V. & Boissin, E. On the occurrence of intracolonial genotypic variability in highly clonal populations of the hydrocoral Millepora platyphylla at Moorea (French Polynesia). Sci. Rep. 7, 1–10. https://doi.org/10.1038/s41598-017-14684-3 (2017).CAS 
    Article 

    Google Scholar 
    20.Maier, E., Buckenmaier, A., Tollrian, R. & Nürnberger, B. Intracolonial genetic variation in the scleractinian coral Seriatopora hystrix. Coral Reefs 31, 505–517. https://doi.org/10.1007/s00338-011-0857-9 (2012).ADS 
    Article 

    Google Scholar 
    21.Schweinsberg, M., Weiss, L. C., Striewski, S., Tollrian, R. & Lampert, K. P. More than one genotype: How common is intracolonial genetic variability in scleractinian corals? Mol. Ecol. 24, 2673–2685. https://doi.org/10.1111/mec.13200 (2015).Article 
    PubMed 

    Google Scholar 
    22.van Oppen, M. J., Souter, P., Howells, E. J., Heyward, A. & Berkelmans, R. Novel genetic diversity through somatic mutations: Fuel for adaptation of reef corals? Diversity 3, 405–423 (2011).Article 

    Google Scholar 
    23.Rinkevich, B. A critical approach to the definition of Darwinian units of selection. Biol. Bull. 199, 231–240. https://doi.org/10.2307/1543179 (2000).CAS 
    Article 
    PubMed 

    Google Scholar 
    24.Rinkevich, B. The apex set-up for the major transitions in individuality. Evol. Biol. 46, 217–228. https://doi.org/10.1007/s11692-019-09481-x (2019).Article 

    Google Scholar 
    25.Santelices, B. How many kinds of individual are there? Trends Ecol. Evol. 14, 152–155. https://doi.org/10.1016/S0169-5347(98)01519-5 (1999).CAS 
    Article 
    PubMed 

    Google Scholar 
    26.Pineda-Krch, M. & Lehtilä, K. Costs and benefits of genetic heterogeneity within organisms. J. Evol. Biol. 17, 1167–1177. https://doi.org/10.1111/j.1420-9101.2004.00808.x (2004).CAS 
    Article 
    PubMed 

    Google Scholar 
    27.Rinkevich, B. Quo vadis chimerism? Chimerism 2, 1–5. https://doi.org/10.4161/chim.14725 (2011).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    28.Rinkevich, B. & Yankelevich, I. Environmental split between germ cell parasitism and somatic cell synergism in chimeras of a colonial urochordate. J. Exp. Biol. 207, 3531–3536. https://doi.org/10.1242/jeb.01184 (2004).Article 
    PubMed 

    Google Scholar 
    29.Raymundo, L. J. & Maypa, A. P. Getting bigger faster: Mediation of size-specific mortality via fusion in juvenile coral transplants. Ecol. Appl. 14, 281–295. https://doi.org/10.1890/02-5373 (2004).Article 

    Google Scholar 
    30.Rinkevich, B., Shaish, L., Douek, J. & Ben-Shlomo, R. Venturing in coral larval chimerism: A compact functional domain with fostered genotypic diversity. Sci. Rep. 6, 19493. https://doi.org/10.1038/srep19493 (2016).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    31.Rinkevich, B. Coral chimerism as an evolutionary rescue mechanism to mitigate global climate change impacts. Glob. Chang Biol. 25, 1198–1206. https://doi.org/10.1111/gcb.14576 (2019).ADS 
    Article 

    Google Scholar 
    32.Amar, K.-O., Chadwick, N. E. & Rinkevich, B. Coral planulae as dispersion vehicles: Biological properties of larvae released early and late in the season. Mar. Ecol. Prog. Ser. 350, 71–78. https://doi.org/10.3354/meps07125 (2007).ADS 
    Article 

    Google Scholar 
    33.Rinkevich, B. Immunology of human implantation: From the invertebrate’s point of view. Hum. Reprod. 13, 455–459. https://doi.org/10.1093/humrep/13.2.503 (1998).CAS 
    Article 
    PubMed 

    Google Scholar 
    34.González, A. V. & Santelices, B. Frequency of chimerism in populations of the kelp Lessonia spicata in central Chile. PLoS One 12, 1–20. https://doi.org/10.1371/journal.pone.0169182 (2017).CAS 
    Article 

    Google Scholar 
    35.Nozawa, Y. & Hirose, M. When does the window close? The onset of allogeneic fusion 2–3 years post-settlement in the scleractinian coral, Echinophyllia aspera. Zool. Stud. 50, 396 (2011).
    Google Scholar 
    36.Puill-Stephan, E., van Oppen, M. J. H., Pichavant-Rafini, K. & Willis, B. L. High potential for formation and persistence of chimeras following aggregated larval settlement in the broadcast spawning coral, Acropora millepora. Proc. R. Soc. Lond. B Biol. Sci. 279, 699–708. https://doi.org/10.1098/rspb.2011.1035 (2012).CAS 
    Article 

    Google Scholar 
    37.Frank, U., Oren, U., Loya, Y. & Rinkevich, B. Alloimmune maturation in the coral Stylophora pistillata is achieved through three distinctive stages, 4 months post-metamorphosis. Proc. R. Soc. Lond. B Biol. Sci. 264, 99–104. https://doi.org/10.1098/rspb.1997.0015 (1997).ADS 
    Article 

    Google Scholar 
    38.Rinkevich, B. The branching coral Stylophora pistillata: Contribution of genetics in shaping colony landscape. Isr. J. Zool. 48, 71–82. https://doi.org/10.1560/BCPA-UM3A-MKBP-HGL2 (2002).Article 

    Google Scholar 
    39.Highsmith, R. Reproduction by fragmentation in corals. Mar. Ecol. Prog. Ser. 7, 207–226. https://doi.org/10.3354/meps007207 (1982).ADS 
    Article 

    Google Scholar 
    40.Barfield, S., Aglyamova, G. V. & Matz, M. V. Evolutionary origins of germline segregation in Metazoa: Evidence for a germ stem cell lineage in the coral Orbicella faveolata (Cnidaria, Anthozoa). Proc. R. Soc. Lond. B Biol. Sci. 283, 20152128. https://doi.org/10.1098/rspb.2015.2128 (2016).CAS 
    Article 

    Google Scholar 
    41.Chang, E. S., Orive, M. E. & Cartwright, P. Nonclonal coloniality: Genetically chimeric colonies through fusion of sexually produced polyps in the hydrozoan Ectopleura larynx. Evol. Lett. 2, 442–455. https://doi.org/10.1002/evl3.68 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    42.Hancock, J. P., Goulden, N. J., Oakhill, A. & Steward, C. G. Quantitative analysis of chimerism after allogeneic bone marrow transplantation using immunomagnetic selection and fluorescent microsatellite PCR. Leukemia 17, 247–251. https://doi.org/10.1038/sj.leu.2402759 (2003).CAS 
    Article 
    PubMed 

    Google Scholar 
    43.Broestl, L., Rubin, J. B. & Dahiya, S. Fetal microchimerism in human brain tumors. Brain Pathol. 28, 484–494. https://doi.org/10.1111/bpa.12557 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    44.Olsen, K. C., Moscoso, J. A. & Levitan, D. R. Somatic mutation is a function of clone size and depth in orbicella reef-building corals. Biol. Bull. 236, 1–12. https://doi.org/10.1086/700261 (2019).CAS 
    Article 
    PubMed 

    Google Scholar 
    45.Schweinsberg, M., Tollrian, R. & Lampert, K. P. Inter- and intra-colonial genotypic diversity in hermatypic hydrozoans of the family Milleporidae. Mar. Ecol. 38, 1–11. https://doi.org/10.1111/maec.12388 (2017).Article 

    Google Scholar 
    46.Santelices, B., Alvarado, J. L. & Flores, V. Size increments due to interindividual fusions: How much and for how long? J. Phycol. 46, 685–692. https://doi.org/10.1111/j.1529-8817.2010.00864.x (2010).Article 

    Google Scholar 
    47.Rinkevich, B. & Weissman, I. L. Chimeras vs genetically homegeneous individuals: Potential fitness costs and benefits. Oikos 63, 119–124 (1992).Article 

    Google Scholar 
    48.Mizrahi, D., Navarrete, S. A. & Flores, A. A. V. Groups travel further: Pelagic metamorphosis and polyp clustering allow higher dispersal potential in sun coral propagules. Coral Reefs 33, 443–448. https://doi.org/10.1007/s00338-014-1135-4 (2014).ADS 
    Article 

    Google Scholar 
    49.Lambert, N. C. et al. Quantification of maternal microchimerism by HLA-specific real-time polymerase chain reaction: Studies of healthy women and women with scleroderma. Arthritis Rheumatol. 50, 906–914. https://doi.org/10.1002/art.20200 (2004).CAS 
    Article 

    Google Scholar 
    50.Magor, B. G., De Tomoso, A., Rinkevich, B. & Weissman, I. L. Allorecognition in colonial tunicates: Protection against predatory cell lineages? Immunol. Rev. 167, 69–79. https://doi.org/10.1111/j.1600-065x.1999.tb01383.x (1999).CAS 
    Article 
    PubMed 

    Google Scholar 
    51.Duerden, J. E. Aggregated colonies in madreporarian corals. Am. Nat. 34, 461–471 (1902).Article 

    Google Scholar 
    52.Barki, Y., Gateño, D., Graur, D. & Rinkevich, B. Soft-coral natural chimerism: A window in ontogeny allows the creation of entities comprised of incongruous parts. Mar. Ecol. Prog. Ser. 231, 91–99. https://doi.org/10.3354/meps231091 (2002).ADS 
    Article 

    Google Scholar 
    53.Linden, B., Huisman, J. & Rinkevich, B. Circatrigintan instead of lunar periodicity of larval release in a brooding coral species. Sci. Rep. 8, 5668. https://doi.org/10.1038/s41598-018-23274-w (2018).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    54.Shefy, D., Shashar, N. & Rinkevich, B. The reproduction of the Red Sea coral Stylophora pistillata from Eilat: 4-decade perspective. Mar. Biol. 165, 27. https://doi.org/10.1007/s00227-017-3280-0 (2018).Article 

    Google Scholar 
    55.Shafir, S., Van Rijn, J. & Rinkevich, B. Steps in the construction of underwater coral nursery, an essential component in reef restoration acts. Mar. Biol. 149, 679–687. https://doi.org/10.1007/s00227-005-0236-6 (2006).Article 

    Google Scholar 
    56.Rinkevich, B. & Loya, Y. The reproduction of the Red Sea coral Stylophora pistillata. I. Gonads and planulae. Mar. Ecol. Prog. Ser. 1, 133–144 (1979).ADS 
    Article 

    Google Scholar 
    57.Santelices, B. Mosaicism and chimerism as components of intraorganismal genetic heterogeneity. J. Evol. Biol. 17, 1187–1188. https://doi.org/10.1111/j.1420-9101.2004.00813.x (2004).CAS 
    Article 
    PubMed 

    Google Scholar 
    58.Graham, D. E. The isolation of high molecular weight DNA from whole organisms or large tissue masses. Anal. Biochem. 85, 609–613. https://doi.org/10.1016/0003-2697(78)90262-2 (1978).CAS 
    Article 
    PubMed 

    Google Scholar 
    59.Douek, J., Barki, Y., Gateño, D. & Rinkevich, B. Possible cryptic speciation within the sea anemone Actinia equina complex detected by AFLP markers. Zool. J. Linn. Soc. 136, 315–320. https://doi.org/10.1046/j.1096-3642.2002.00034.x (2002).Article 

    Google Scholar 
    60.Banguera-Hinestroza, E., Saenz-Agudelo, P., Bayer, T., Berumen, M. L. & Voolstra, C. R. Characterization of new microsatellite loci for population genetic studies in the smooth cauliflower coral (Stylophora sp.). Conserv. Genet. Resour. 5, 561–563. https://doi.org/10.1007/s12686-012-9852-x (2013).Article 

    Google Scholar 
    61.Diwan, N. & Cregan, P. B. Automated sizing of fluorescent-labeled simple sequence repeat (SSR) markers to assay genetic variation in soybean. Theor. Appl. Genet. 95, 723–733. https://doi.org/10.1007/s001220050618 (1997).CAS 
    Article 

    Google Scholar 
    62.Hearne, C. M., Ghosh, S. & Todd, J. A. Microsatellites for linkage analysis of genetic traits. Trends Genet. 8, 288–294. https://doi.org/10.1016/0168-9525(92)90256-4 (1992).CAS 
    Article 
    PubMed 

    Google Scholar  More