More stories

  • in

    Environmental optima for an ecosystem engineer: a multidisciplinary trait-based approach

    1.Rivadeneira, M. M. et al. Testing the abundant-centre hypothesis using intertidal porcelain crabs along the Chilean coast: Linking abundance and life-history variation. J. Biogeogr. 37, 486–498 (2010).
    Google Scholar 
    2.Hutchins, L. W. The bases for temperature zonation in geographical distribution. Ecol. Monogr. 17, 325–335 (1947).
    Google Scholar 
    3.Lewis, J. R. Latitudinal trends in reproduction, recruitment and population characteristics of some rocky littoral molluscs and cirripedes. Hydrobiologia 142, 1–13 (1986).
    Google Scholar 
    4.Bernardo, J. The particular maternal effect of propagule size, especially egg size: Patterns, models, quality of evidence and interpretations. Am. Zool. 36, 216–236 (1996).
    Google Scholar 
    5.Thorson, G. Reproductive and larval ecology of marine bottom invertebrates. Biol. Rev. 25, 1–45 (1950).CAS 
    PubMed 

    Google Scholar 
    6.Marshall, D. J., Pettersen, A. K. & Cameron, H. A global synthesis of offspring size variation, its eco-evolutionary causes and consequences. Funct. Ecol. 32, 1436–1446 (2018).
    Google Scholar 
    7.Des Roches, S. et al. The ecological importance of intraspecific variation. Nat. Ecol. Evol. 2, 57–64 (2018).PubMed 

    Google Scholar 
    8.Violle, C. et al. Let the concept of trait be functional!. Oikos 116, 882–892 (2007).
    Google Scholar 
    9.Sides, C. B. et al. Revisiting Darwin’s hypothesis: Does greater intraspecific variability increase species’ ecological breadth?. Am. J. Bot. 101, 56–62 (2014).PubMed 

    Google Scholar 
    10.Moran, E. V., Hartig, F. & Bell, D. M. Intraspecific trait variation across scales: Implications for understanding global change responses. Glob. Change Biol. 22, 137–150 (2016).ADS 

    Google Scholar 
    11.Violle, C. et al. The return of the variance: Intraspecific variability in community ecology. Trends Ecol. Evol. 27, 244–252 (2012).PubMed 

    Google Scholar 
    12.Stark, J., Lehman, R., Crawford, L., Enquist, B. J. & Blonder, B. Does environmental heterogeneity drive functional trait variation? A test in montane and alpine meadows. Oikos 126, 1650–1659 (2017).
    Google Scholar 
    13.Stearns, S. C. The Evolution of Life Histories. xii, 249p. No. 575 S81 (Oxford, Oxford University, 1992).14.Vance, R. R. On reproductive strategies in marine benthic invertebrates. Am. Nat. 107, 339–352 (1973).
    Google Scholar 
    15.Levitan, D. R. Gamete traits influence the variance in reproductive success, the intensity of sexual selection, and the outcome of sexual conflict among congeneric sea urchins. Evolution 62, 1305–1316 (2008).PubMed 

    Google Scholar 
    16.Lavorel, S. & Garnier, E. Predicting changes in community composition and ecosystem functioning from plant traits: revisiting the Holy Grail. Funct. Ecol. 16, 545–556 (2002).
    Google Scholar 
    17.Pineda, M. C. et al. Tough adults, frail babies: An analysis of stress sensitivity across early life-history stages of widely introduced marine invertebrates. PLoS ONE 7, e46672 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    18.Harley, C. D. G. et al. The impacts of climate change in coastal marine systems: Climate change in coastal marine systems. Ecol. Lett. 9, 228–241 (2006).ADS 
    PubMed 

    Google Scholar 
    19.Stein, A., Gerstner, K. & Kreft, H. Environmental heterogeneity as a universal driver of species richness across taxa, biomes and spatial scales. Ecol. Lett. 17, 866–880 (2014).PubMed 

    Google Scholar 
    20.Foo, S. A. & Byrne, M. Marine gametes in a changing ocean: Impacts of climate change stressors on fecundity and the egg. Mar. Environ. Res. 128, 12–24 (2017).CAS 
    PubMed 

    Google Scholar 
    21.Dahlhoff, E. P. Biochemical indicators of stress and metabolism: Applications for marine ecological studies. Annu. Rev. Physiol. 66, 183–207 (2004).CAS 
    PubMed 

    Google Scholar 
    22.Soudant, P. et al. Comparison of the lipid class and fatty acid composition between a reproductive cycle in nature and a standard hatchery conditioning of the Pacific Oyster Crassostrea gigas. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 123, 209–222 (1999).
    Google Scholar 
    23.Lester, S. E., Gaines, S. D. & Kinlan, B. P. Reproduction on the edge: Large-scale patterns of individual performance in a marine invertebrate. Ecology 88, 2229–2239 (2007).PubMed 

    Google Scholar 
    24.Helmuth, B., Mieszkowska, N., Moore, P. & Hawkins, S. J. Living on the edge of two changing worlds: Forecasting the responses of rocky intertidal ecosystems to climate change. Annu. Rev. Ecol. Evol. Syst. 37, 373–404 (2006).
    Google Scholar 
    25.Sagarin, R. D., Barry, J. P., Gilman, S. E. & Baxter, C. H. Climate-related change in an intertidal community over short and long time scales. Ecol. Monogr. 69, 465–490 (1999).
    Google Scholar 
    26.Dubois, S., Retière, C. & Olivier, F. Biodiversity associated with Sabellaria alveolata (Polychaeta: Sabellariidae) reefs: Effects of human disturbances. J. Mar. Biol. Assoc. UK 82, 817–826 (2002).
    Google Scholar 
    27.Jones, A. G., Dubois, S. F., Desroy, N. & Fournier, J. Interplay between abiotic factors and species assemblages mediated by the ecosystem engineer Sabellaria alveolata (Annelida: Polychaeta). Estuar. Coast. Shelf Sci. 200, 1–18 (2018).ADS 

    Google Scholar 
    28.Bonifazi, A. et al. Macrofaunal biodiversity associated with different developmental phases of a threatened Mediterranean Sabellaria alveolata (Linnaeus, 1767) reef. Mar. Environ. Res. 145, 97–111 (2019).CAS 
    PubMed 

    Google Scholar 
    29.Holt, T. J., Biogenic Reefs. An Overview of Dynamic and Sensitivity Characteristics for Conservation Management of Marine SACs. UK Marine SACs Project (1998).30.Crisp, D. The effects of the severe winter of 1962–1963 on marine life in Britain. J. Anim. Ecol. 33, 165–210 (1964).
    Google Scholar 
    31.Firth, L. B. et al. Historical comparisons reveal multiple drivers of decadal change of an ecosystem engineer at the range edge. Ecol. Evol. 5, 3210–3222 (2015).PubMed 
    PubMed Central 

    Google Scholar 
    32.Firth, L. B. et al. Specific niche requirements underpin multidecadal range edge stability, but may introduce barriers for climate change adaptation. Divers. Distrib. 27, 668–683 (2021).
    Google Scholar 
    33.Wethey, D. S. et al. Response of intertidal populations to climate: Effects of extreme events versus long term change. J. Exp. Mar. Biol. Ecol. 400, 132–144 (2011).
    Google Scholar 
    34.Sagarin, R. D. & Gaines, S. D. Geographical abundance distributions of coastal invertebrates: Using one-dimensional ranges to test biogeographic hypotheses. J. Biogeogr. 29, 985–997 (2002).
    Google Scholar 
    35.Rahman, M. A., Henderson, S., Miller-Ezzy, P., Li, X. X. & Qin, J. G. Immune response to temperature stress in three bivalve species: Pacific oyster Crassostrea gigas, Mediterranean mussel Mytilus galloprovincialis and mud cockle Katelysia rhytiphora. Fish Shellfish Immunol. 86, 868–874 (2019).CAS 
    PubMed 

    Google Scholar 
    36.Osada, M., Nishikawa, M. & Nomura, T. Involvement of prostaglandins in the spawning of the scallop, Patinopecten yessoensis. Comp. Biochem. Physiol. C 94, 595–601 (1989).
    Google Scholar 
    37.Stanley, D. W. & Howard, R. W. The biology of prostaglandins and related eicosanoids in invertebrates: Cellular, organismal and ecological actions. Am. Zool. 38, 369–381 (1998).CAS 

    Google Scholar 
    38.Pernet, F., Tremblay, R., Comeau, L. & Guderley, H. Temperature adaptation in two bivalve species from different thermal habitats: Energetics and remodelling of membrane lipids. J. Exp. Biol. 210, 2999–3014 (2007).PubMed 

    Google Scholar 
    39.Muir, A. P., Nunes, F. L. D., Dubois, S. F. & Pernet, F. Lipid remodelling in the reef-building honeycomb worm, Sabellaria alveolata, reflects acclimation and local adaptation to temperature. Sci. Rep. 6, 35669 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    40.Hulbert, A. & Else, P. L. Membranes as possible pacemakers of metabolism. J. Theor. Biol. 199, 257–274 (1999).ADS 
    CAS 
    PubMed 

    Google Scholar 
    41.Brokordt, K. B., Himmelman, J. H., Nusetti, O. A. & Guderley, H. E. Reproductive investment reduces recuperation from exhaustive escape activity in the tropical scallop Euvola zizac. Mar. Biol. 137, 857–865 (2000).CAS 

    Google Scholar 
    42.Levitan, D. R. & Roitberg, B. D. Optimal egg size in marine invertebrates: Theory and phylogenetic analysis of the critical relationship between egg size and development time in echinoids. Am. Nat. 156, 175–192 (2000).PubMed 

    Google Scholar 
    43.Moran, A. L. & McAlister, J. S. Egg size as a life history character of marine invertebrates: Is it all it’s cracked up to be?. Biol. Bull. 216, 226–242 (2009).PubMed 

    Google Scholar 
    44.Marshall, D. J. & Burgess, S. C. Deconstructing environmental predictability: Seasonality, environmental colour and the biogeography of marine life histories. Ecol. Lett. 18, 174–181 (2015).PubMed 

    Google Scholar 
    45.Racault, M.-F., Le Quéré, C., Buitenhuis, E., Sathyendranath, S. & Platt, T. Phytoplankton phenology in the global ocean. Ecol. Indic. 14, 152–163 (2012).
    Google Scholar 
    46.Henson, S., Cole, H., Beaulieu, C. & Yool, A. The impact of global warming on seasonality of ocean primary production. Biogeosciences 10, 4357–4369 (2013).ADS 

    Google Scholar 
    47.Morim, J. et al. Robustness and uncertainties in global multivariate wind-wave climate projections. Nat. Clim. Change 9, 711–718 (2019).ADS 

    Google Scholar 
    48.Stillman, J. H. Heat waves, the new normal: Summertime temperature extremes will impact animals, ecosystems, and human communities. Physiology 34, 86–100 (2019).CAS 
    PubMed 

    Google Scholar 
    49.McCarthy, D., Young, C. & Emson, R. Influence of wave-induced disturbance on seasonal spawning patterns in the sabellariid polychaete Phragmatopoma lapidosa. Mar. Ecol. Prog. Ser. 256, 123–133 (2003).ADS 

    Google Scholar 
    50.Aviz, D., Pinto, A. J. A., Ferreira, M. A. P., Rocha, R. M. & Rosa Filho, J. S. Reproductive biology of Sabellaria wilsoni (Sabellariidae: Polychaeta), an important ecosystem engineer on the Amazon coast. J. Mar. Biol. Assoc. UK https://doi.org/10.1017/S0025315416001776 (2016).Article 

    Google Scholar 
    51.Bowman, R. S. & Lewis, J. Annual fluctuations in the recruitment of Patella vulgata L. J. Mar. Biol. Assoc. U. K. 57, 793–815 (1977).
    Google Scholar 
    52.Sagarin, R. D. & Somero, G. N. Complex patterns of expression of heat-shock protein 70 across the southern biogeographical ranges of the intertidal mussel Mytilus californianus and snail Nucella ostrina. J. Biogeogr. 33, 622–630 (2006).
    Google Scholar 
    53.Wernberg, T. et al. An extreme climatic event alters marine ecosystem structure in a global biodiversity hotspot. Nat. Clim. Change 3, 78–82 (2013).ADS 

    Google Scholar 
    54.Firth, L. B., Knights, A. M. & Bell, S. S. Air temperature and winter mortality: Implications for the persistence of the invasive mussel, Perna viridis in the intertidal zone of the south-eastern United States. J. Exp. Mar. Biol. Ecol. 400, 250–256 (2011).
    Google Scholar 
    55.Seabra, R., Wethey, D. S., Santos, A. M. & Lima, F. P. Side matters: Microhabitat influence on intertidal heat stress over a large geographical scale. J. Exp. Mar. Biol. Ecol. 400, 200–208 (2011).
    Google Scholar 
    56.Meneghesso, C. et al. Remotely-sensed L4 SST underestimates the thermal fingerprint of coastal upwelling. Remote Sens. Environ. 237, 111588 (2020).ADS 

    Google Scholar 
    57.Marshall, D. J. & Keough, M. J. The evolutionary ecology of offspring size in marine invertebrates. in Advances in Marine Biology, 1–60. https://doi.org/10.1016/S0065-2881(07)53001-4 (Elsevier, 2007).58.Albert, C. H. et al. A multi-trait approach reveals the structure and the relative importance of intra- vs. interspecific variability in plant traits: Intra- vs. interspecific variability in plant traits. Funct. Ecol. 24, 1192–1201 (2010).
    Google Scholar 
    59.Olofsson, H., Ripa, J. & Jonzén, N. Bet-hedging as an evolutionary game: The trade-off between egg size and number. Proc. R. Soc. B Biol. Sci. 276, 2963–2969 (2009).
    Google Scholar 
    60.Osovitz, C. J. & Hofmann, G. E. Marine macrophysiology: Studying physiological variation across large spatial scales in marine systems. Comp. Biochem. Physiol. A. Mol. Integr. Physiol. 147, 821–827 (2007).PubMed 

    Google Scholar 
    61.Clarke, A. Reproduction in the cold: Thorson revisited. Invertebr. Reprod. Dev. 22, 175–183 (1992).
    Google Scholar 
    62.Hawkins, S. J. et al. Distinguishing globally-driven changes from regional- and local-scale impacts: The case for long-term and broad-scale studies of recovery from pollution. Mar. Pollut. Bull. 124, 573–586 (2017).CAS 
    PubMed 

    Google Scholar 
    63.Dahlhoff, E. P., Stillman, J. H. & Menge, B. A. Physiological community ecology: Variation in metabolic activity of ecologically important rocky intertidal invertebrates along environmental gradients. Integr. Comp. Biol. 42, 862–871 (2002).PubMed 

    Google Scholar 
    64.Nunes, F. L. D., Rigal, F., Dubois, S. F. & Viard, F. Looking for diversity in all the right places? Genetic diversity is highest in peripheral populations of the reef-building polychaete Sabellaria alveolata. Mar. Biol. 168, 63 (2021).
    Google Scholar 
    65.Bush, L. E. Stability and Variability of the Ecosystem Engineer Sabellaria alveolata on Differing Temporal and Spatial Scales (Bangor University, 2016).
    Google Scholar 
    66.Lourenço, C. R., Nicastro, K. R., McQuaid, C. D., Krug, L. A. & Zardi, G. I. Strong upwelling conditions drive differences in species abundance and community composition along the Atlantic coasts of Morocco and Western Sahara. Mar. Biodivers. 50, 15 (2020).
    Google Scholar 
    67.Ritchie, H. & Marshall, D. J. Fertilisation is not a new beginning: Sperm environment affects offspring developmental success. J. Exp. Biol. 216, 3104–3109 (2013).PubMed 

    Google Scholar 
    68.Dubois, S., Comtet, T., Retière, C. & Thiébaut, E. Distribution and retention of Sabellaria alveolata larvae (Polychaeta: Sabellariidae) in the Bay of Mont-Saint-Michel, France. Mar. Ecol. Prog. Ser. 346, 243–254 (2007).ADS 
    CAS 

    Google Scholar 
    69.Costello, D. P., Henley, C., & Marine Biological Laboratory (Woods Hole, Mass.). Methods for obtaining and handling marine eggs and embryos [by] Donald P. Costello and Catherine Henley. ([s.n.], 1971). https://doi.org/10.5962/bhl.title.1020.70.Gruet, Y. Aspects morphologiques et dynamiques de constructions de l’Annélide polychete Sabellaria alveolata (Linne). Rev. Trav. Inst. Pêch. Marit. 36, 131–161 (1972).
    Google Scholar 
    71.Saulquin, B., Gohin, F. & Garrello, R. Regional objective analysis for merging high-resolution MERIS, MODIS/Aqua, and SeaWiFS Chlorophyll-a data from 1998 to 2008 on the European Atlantic Shelf. IEEE Trans. Geosci. Remote Sens. 49, 143–154 (2011).ADS 

    Google Scholar 
    72.Gohin, F. Annual cycles of chlorophyll-a, non-algal suspended particulate matter, and turbidity observed from space and in-situ in coastal waters. Ocean Sci. 7, 705–732 (2011).ADS 
    CAS 

    Google Scholar 
    73.Seabra, R., Wethey, D. S., Santos, A. M. & Lima, F. P. Understanding complex biogeographic responses to climate change. Sci. Rep. 5, 12930 (2015).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    74.Schlegel, R. W., Darmaraki, S., Benthuysen, J. A., Filbee-Dexter, K. & Oliver, E. C. J. Marine cold-spells. Progress Oceanogr. 198, 102684 (2021).
    Google Scholar 
    75.Hobday, A. J. et al. A hierarchical approach to defining marine heatwaves. Prog. Oceanogr. 141, 227–238 (2016).ADS 

    Google Scholar 
    76.Schlegel, R. W., Oliver, E. C., Hobday, A. J. & Smit, A. J. Detecting marine heatwaves with sub-optimal data. Front. Mar. Sci. 6, 737 (2019).
    Google Scholar 
    77.Egbert, G. D., Erofeeva, S. Y. & Ray, R. D. Assimilation of altimetry data for nonlinear shallow-water tides: Quarter-diurnal tides of the Northwest European Shelf. Cont. Shelf Res. 30, 668–679 (2010).ADS 

    Google Scholar 
    78.Burrows, M., Harvey, R. & Robb, L. Wave exposure indices from digital coastlines and the prediction of rocky shore community structure. Mar. Ecol. Prog. Ser. 353, 1–12 (2008).ADS 

    Google Scholar 
    79.Wessel, P. & Smith, W. H. F. A global, self-consistent, hierarchical, high-resolution shoreline database. J. Geophys. Res. Solid Earth 101, 8741–8743 (1996).
    Google Scholar 
    80.Seers, B. fetchR: Calculate Wind Fetch. R Package Version 2-1 (2017).81.Guillaume, A. S., Monro, K. & Marshall, D. J. Transgenerational plasticity and environmental stress: Do paternal effects act as a conduit or a buffer?. Funct. Ecol. 30, 1175–1184 (2016).
    Google Scholar 
    82.Curd, A. et al. Connecting organic to mineral: How the physiological state of an ecosystem-engineer is linked to its habitat structure. Ecol. Indic. 98, 49–60 (2019).CAS 

    Google Scholar 
    83.Gruet, Y. & Lassus, P. Contribution a l’etude de la biologie reproductive d’une population naturelle de l’Annelide Polychete, Sabellaria alveolata (Linnaeus). Ann. Inst. Oceanogr. Monaco 59, 127–140 (1983).
    Google Scholar 
    84.Hazel, J. The role of alterations in membrane lipid composition in enabling physiological adaptation of organisms to their physical environment. Prog. Lipid Res. 29, 167–227 (1990).CAS 
    PubMed 

    Google Scholar 
    85.Hochachka, P. W. & Somero, G. N. Biochemical Adaptation: Mechanism and Process in Physiological Evolution (Oxford University Press, 2002).
    Google Scholar 
    86.Abele, D. & Puntarulo, S. Formation of reactive species and induction of antioxidant defence systems in polar and temperate marine invertebrates and fish. Comp. Biochem. Physiol. A Mol. Integr. Physiol. 138, 405–415 (2004).PubMed 

    Google Scholar 
    87.Folch, J., Lees, M. & Stanley, G. H. S. A simple method for the isolation and purification of total lipides from animal tissues. J. Biol. Chem. 226, 497–509. http://www.jbc.org/content/226/1/497 (1957).CAS 
    PubMed 

    Google Scholar 
    88.Sieracki, C., Sieracki, M. & Yentsch, C. An imaging-in-flow system for automated analysis of marine microplankton. Mar. Ecol. Prog. Ser. 168, 285–296 (1998).ADS 

    Google Scholar 
    89.Pasteels, J. J. Etude au microscope électronique de la réaction corticale. II. La réaction corticale de l’oeuf vierge de Sabellaria alveolata. J. Embryol. Exp. Morphol. 13, 327–339 (1965).CAS 
    PubMed 

    Google Scholar 
    90.Doledec, S. & Chessel, D. Co-inertia analysis: An alternative method for studying species-environment relationships. Freshw. Biol. 31, 277–294 (1994).
    Google Scholar 
    91.Robert, P. & Escoufier, Y. A unifying tool for linear multivariate statistical methods: The RV-coefficient. J. R. Stat. Soc. Ser. C Appl. Stat. 25, 257–265 (1976).MathSciNet 

    Google Scholar 
    92.Legendre, P. & Legendre, L. Ecological resemblance. in Developments in Environmental Modelling Chapter 7, Vol. 24, 265–335 (Elsevier, 2012).93.Borcard, D., Legendre, P. & Drapeau, P. Partialling out the spatial component of ecological variation. Ecology 73, 1045–1055 (1992).
    Google Scholar 
    94.Peres-Neto, P. R., Legendre, P., Dray, S. & Borcard, D. Variation partitioning of species data matrices: Estimation and comparison of fractions. Ecology 87, 2614–2625 (2006).PubMed 

    Google Scholar 
    95.Dormann, C. F. et al. Collinearity: A review of methods to deal with it and a simulation study evaluating their performance. Ecography 36, 27–46 (2013).
    Google Scholar 
    96.Messier, J., McGill, B. J. & Lechowicz, M. J. How do traits vary across ecological scales? A case for trait-based ecology: How do traits vary across ecological scales?. Ecol. Lett. 13, 838–848 (2010).PubMed 

    Google Scholar 
    97.Rao, C. R. The use and interpretation of principal component analysis in applied research. Sankhyā Indian J. Stat. Ser. A (1961-2002) 26, 329–358 (1964).98.R Core Team: A language and environment for statistical computing. Available from: https://www.R-project.org/ (2018).99.Oksanen, J. et al. Package ‘vegan’. Commun. Ecol. Package Version 2, 1–295 (2013).
    Google Scholar  More

  • in

    A database of global coastal conditions

    1.Horning, N., Robinson, J. A., Sterling, E. J., Turner, W. & Spector, S. Remote sensing for ecology and conservation. Techniques in Ecology & Conservation Series (Oxford University Press, 2010).2.Li, J. et al. A review of remote sensing for environmental monitoring in China. Remote Sens. 12, 1130 (2020).ADS 

    Google Scholar 
    3.Carter, W. D. & Paulson, R. W. Introduction to monitoring dynamic environmental phenomena of the world using satellite data collection systems. (U.S. Geological Survey, 1979).4.Nurdin, S., Mustapha, M. A. & Lihan, T. The relationship between sea surface temperature and chlorophyll-a concentration in fisheries aggregation area in the archipelagic waters of spermonde using satellite images. AIP Conf. Proc. 1571, 466–472 (2013).ADS 

    Google Scholar 
    5.Ward, D., Phinn, S. R. & Murray, A. T. Monitoring growth in rapidly urbanizing areas using remotely sensed data. Prof. Geogr. 52, 371–386 (2000).
    Google Scholar 
    6.Singh, A. Review article: Digital change detection techniques using remotely-sensed data. Int. J. Remote Sens. 10, 989–1003 (1989).
    Google Scholar 
    7.Dewan, A. M. & Yamaguchi, Y. Land use and land cover change in Greater Dhaka, Bangladesh: Using remote sensing to promote sustainable urbanization. Appl. Geogr. 29, 390–401 (2009).
    Google Scholar 
    8.Green, K., Kempka, D. & Lackey, L. Using remote sensing to detect and monitor land-cover and land-use change. Photogramm. Eng. Remote Sens. 60, 331–337 (1994).
    Google Scholar 
    9.Nagendra, H. Using remote sensing to assess biodiversity. Int. J. Remote Sens. 22, 2377–2400 (2001).
    Google Scholar 
    10.Rosenqvist, Å., Milne, A., Lucas, R., Imhoff, M. & Dobson, C. A review of remote sensing technology in support of the Kyoto Protocol. Environ. Sci. Policy 6, 441–455 (2003).
    Google Scholar 
    11.Liu, J. A process-based boreal ecosystem productivity simulator using remote sensing inputs. Remote Sens. Environ. 62, 158–175 (1997).ADS 

    Google Scholar 
    12.Colwell, R. R. Global climate and infectious disease: The cholera paradigm. Science 274, 2025–2031 (1996).ADS 
    PubMed 
    CAS 

    Google Scholar 
    13.Escobar, L. E. et al. A global map of suitability for coastal Vibrio cholerae under current and future climate conditions. Acta Trop. 149, 202–211 (2015).PubMed 

    Google Scholar 
    14.Watts, N. et al. The 2019 report of The Lancet Countdown on health and climate change: Ensuring that the health of a child born today is not defined by a changing climate. Lancet 394, 1836–1878 (2019).PubMed 

    Google Scholar 
    15.Alesheikh, A. A., Ghorbanali, A. & Nouri, N. Coastline change detection using remote sensing. Int. J. Environ. Sci. Technol. 4, 61–66 (2007).
    Google Scholar 
    16.Specter, C. & Gayle, D. Managing technology transfer for coastal zone development: Caribbean experts identify major issues. Int. J. Remote Sens. 11, 1729–1740 (1990).
    Google Scholar 
    17.Green, E. P., Mumby, P. J., Edwards, A. J. & Clark, C. D. A review of remote sensing for the assessment and management of tropical coastal resources. Coast. Manag. 24, 1–40 (1996).
    Google Scholar 
    18.NASA. MODIS (Moderate Resolution Imaging Spectroradiometer). https://modis.gsfc.nasa.gov/about/ (2021).19.Kilpatrick, K. A. et al. A decade of sea surface temperature from MODIS. Remote Sens. Environ. 165, 27–41 (2015).ADS 

    Google Scholar 
    20.Esaias, W. E. et al. An overview of MODIS capabilities for ocean science observations. IEEE Trans. Geosci. Remote Sens. 36, 1250–1265 (1998).ADS 

    Google Scholar 
    21.Donlon, C. J. et al. Toward improved validation of satellite SST measurements for climate research. J. Clim. 15, 353–369 (2002).ADS 

    Google Scholar 
    22.Minnett, P. J. Satellite infrared scanning radiometers — AVHRR and ATSR/M. in Microwave Remote Sensing for Oceanographic and Marine Weather-Forecast Models 141–163 (Springer Netherlands, 1990).23.Hillger, D. et al. First-Light Imagery from Suomi NPP VIIRS. Bull. Am. Meteorol. Soc. 94, 1019–1029 (2013).ADS 

    Google Scholar 
    24.O’Brien, J. From MODIS to VIIRS – Making the Switch for Air Quality Professionals. NASA Earth Science/Applied Science https://appliedsciences.nasa.gov/our-impact/news/modis-viirs-making-switch-air-quality-professionals (2020).25.Minnett, P. J., Evans, R. H., Podestá, G. P. & Kilpatrick, K. A. Sea-surface temperature from Suomi-NPP VIIRS: Algorithm development and uncertainty estimation. in SPIE 9111, Ocean Sensing and Monitoring VI (eds. Hou, W. W. & Arnone, R. A.) 91110C (2014).26.Drusch, M. et al. Sentinel-2: ESA’s Optical High-Resolution Mission for GMES Operational Services. Remote Sens. Environ. 120, 25–36 (2012).ADS 

    Google Scholar 
    27.Donlon, C. et al. The global ocean data assimilation experiment high-resolution sea surface temperature pilot project. Bull. Am. Meteorol. Soc. 88, 1197–1214 (2007).ADS 

    Google Scholar 
    28.NOAA. Ocean Facts: Why do scientists measure sea surface temperature? https://oceanservice.noaa.gov/facts/sea-surface-temperature.html (2020).29.Wei, G. F., Tang, D. L. & Wang, S. Distribution of chlorophyll and harmful algal blooms (HABs): A review on space based studies in the coastal environments of Chinese marginal seas. Adv. Sp. Res. 41, 12–19 (2008).ADS 
    CAS 

    Google Scholar 
    30.O’Reilly, J. E. et al. Ocean color chlorophyll algorithms for SeaWiFS. J. Geophys. Res. Ocean. 103, 24937–24953 (1998).ADS 

    Google Scholar 
    31.Hu, C., Lee, Z. & Franz, B. Chlorophyll a algorithms for oligotrophic oceans: A novel approach based on three-band reflectance difference. J. Geophys. Res. Ocean. 117, C01011 (2012).ADS 

    Google Scholar 
    32.Vezzulli, L. et al. Climate influence on Vibrio and associated human diseases during the past half-century in the coastal North Atlantic. Proc. Natl. Acad. Sci. 113, E5062–E5071 (2016).PubMed 
    PubMed Central 
    CAS 

    Google Scholar 
    33.Lipp, E. K., Huq, A. & Colwell, R. R. Effects of global climate on infectious disease: The Cholera model. Clin. Microbiol. Rev. 15, 757–770 (2002).PubMed 
    PubMed Central 

    Google Scholar 
    34.Grimes, J. D. et al. Viewing marine bacteria, their activity and response to environmental drivers from orbit: Satellite remote sensing of bacteria. Microb. Ecol. 67, 489–500 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    35.Shen, L., Xu, H. & Guo, X. Satellite remote sensing of harmful algal blooms (HABs) and a potential synthesized framework. Sensors 12, 7778–803 (2012).36.Hayashi, M., Jin, F. & Stuecker, M. F. Dynamics for El Niño-La Niña asymmetry constrain equatorial-Pacific warming pattern. Nat. Commun. 11, 1–10 (2020).
    Google Scholar 
    37.Hughes, T. P. et al. Spatial and temporal patterns of mass bleaching of corals in the Anthropocene. Science 359, 80–83 (2018).38.Minnett, P. J. et al. Sea-surface temperature measurements from the moderate-resolution imaging spectroradiometer (MODIS) on Aqua and Terra. in IEEE International Geoscience and Remote Sensing Symposium Proceedings. 2004 7, 4576–4579 (2004).39.Minnett, P. J. The validation of sea surface temperature retrievals from spaceborne infrared radiometers. in Oceanography from Space (Springer Netherlands, 2010).40.Minnett, P. J. & Corlett, G. K. A pathway to generating climate data records of sea-surface temperature from satellite measurements. Deep Sea Res. Part II Top. Stud. Oceanogr. 77–80, 44–51 (2012).ADS 

    Google Scholar 
    41.Castaneda-Guzman, M., Mantilla-Saltos, G., Murray, K. A., Settlage, R. & Escobar, L. E. A database of global coastal conditions. Figshare https://doi.org/10.6084/m9.figshare.c.5660263.v1 (2021).42.R Core Team. R: A Language and Environment for Statistical Computing. (2020).43.NOAA. National Oceanic and Atmospheric Administration (NOAA) Coastal Watch. https://coastwatch.pfeg.noaa.gov/erddapinfo/ (2021).44.Castaneda-Guzman, M., Mantilla-Saltos, G., Murray, K. A., Settlage, R. & Escobar, L. E. Methods and code. Figshare https://doi.org/10.6084/m9.figshare.13708642.v4 (2021).45.Stanford. Best practices for file formats. https://library.stanford.edu/research/data-management-services/data-best-practices/best-practices-file-formats (2021).46.UCAR Community Programs. Network Common Data Form (NetCDF). https://www.unidata.ucar.edu/software/netcdf/ (2021).47.Michna, P. & Woods, M. RNetCDF: Interface to ‘NetCDF’ Datasets. (2019).48.Hijmans, R. J. raster: Geographic Data Analysis and Modeling. (2020).49.ArcGIS. What is a raster data? https://desktop.arcgis.com/en/arcmap/10.3/manage-data/raster-and-images/what-is-raster-data.htm (2021).50.United Nations. United Nations Convention on the Law of the Sea. 1833 U.N.T.S. 397 (1982).51.Tilstone, G. H. et al. Assessment of MODIS-Aqua chlorophyll-a algorithms in coastal and shelf waters of the eastern Arabian Sea. Cont. Shelf Res. 65, 14–26 (2013).ADS 

    Google Scholar 
    52.Hoge, F. E. et al. Validation of Terra-MODIS phytoplankton chlorophyll fluorescence line height. I. Initial airborne Lidar results. Appl. Opt. 42, 2767-2771 (2003).ADS 
    PubMed 

    Google Scholar 
    53.Remer, L. A. Validation of MODIS aerosol retrieval over ocean. Geophys. Res. Lett. 29, 8008 (2002).ADS 

    Google Scholar 
    54.Gentemann, C. L. Three way validation of MODIS and AMSR-E sea surface temperatures. J. Geophys. Res. Ocean. 119, 2583–2598 (2014).ADS 

    Google Scholar 
    55.Fang, H., Wei, S. & Liang, S. Validation of MODIS and CYCLOPES LAI products using global field measurement data. Remote Sens. Environ. 119, 43–54 (2012).ADS 

    Google Scholar 
    56.Hosoda, K., Murakami, H., Sakaida, F. & Kawamura, H. Algorithm and validation of sea surface temperature observation using MODIS sensors aboard terra and aqua in the western North Pacific. J. Oceanogr. 63, 267–280 (2007).
    Google Scholar 
    57.Hao, Y. et al. Validation of MODIS sea surface temperature product in the coastal waters of the Yellow Sea. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 10, 1667–1680 (2017).ADS 

    Google Scholar 
    58.Sims, D. A. et al. On the use of MODIS EVI to assess gross primary productivity of North American ecosystems. J. Geophys. Res. Biogeosciences 111 (2006).59.Miles, T. N. & He, R. Temporal and spatial variability of Chl-a and SST on the South Atlantic Bight: Revisiting with cloud-free reconstructions of MODIS satellite imagery. Cont. Shelf Res. 30, 1951–1962 (2010).ADS 

    Google Scholar 
    60.Ma, S., Zhang, X., Ding, C., Han, W. & Lu, Y. Comparison of the spatiotemporal variation of Chl-a in the East China Sea and Bohai Sea based on long time series satellite data. in 2021 9th International Conference on Agro-Geoinformatics (Agro-Geoinformatics) 1–6 (2021).61.Watts, N. et al. The 2020 report of The Lancet Countdown on health and climate change: Responding to converging crises. Lancet 6736 (2020).62.Moradi, M. & Kabiri, K. Spatio-temporal variability of SST and Chlorophyll-a from MODIS data in the Persian Gulf. Mar. Pollut. Bull. 98, 14–25 (2015).PubMed 
    CAS 

    Google Scholar 
    63.Golder, M. R. et al. Chlorophyll-a, SST and particulate organic carbon in response to the cyclone Amphan in the Bay of Bengal. J. Earth Syst. Sci. 130, 157 (2021).ADS 
    CAS 

    Google Scholar 
    64.Minnett, P. J., Evans, R. H., Kearns, E. J. & Brown, O. B. Sea-surface temperature measured by the Moderate Resolution Imaging Spectroradiometer (MODIS). in IEEE International Geoscience and Remote Sensing Symposium vol. 2, 1177–1179 (IEEE, 2002).65.Qin, H., Chen, G., Wang, W., Wang, D. & Zeng, L. Validation and application of MODIS-derived SST in the South China Sea. Int. J. Remote Sens. 35, 4315–4328 (2014).
    Google Scholar 
    66.Saulquin, B., Gohin, F. & Garrello, R. Regional Objective Analysis for Merging High-Resolution MERIS, MODIS/Aqua, and SeaWiFS Chlorophyll-a Data From 1998 to 2008 on the European Atlantic Shelf. IEEE Trans. Geosci. Remote Sens. 49, 143–154 (2011).ADS 

    Google Scholar 
    67.Chen, J. & Quan, W. An improved algorithm for retrieving chlorophyll-a from the Yellow River Estuary using MODIS imagery. Environ. Monit. Assess. 185, 2243–2255 (2013).PubMed 

    Google Scholar 
    68.Hanafin, J. A. & Minnett, P. J. Thermal profiling of the sea surface skin layer using FTIR measurements. in Gas Transfer at Water Surfaces 161–166 (Blackwell Publishing, 2002).69.Wong, E. W. & Minnett, P. J. The response of the ocean thermal skin layer to variations in incident infrared radiation. J. Geophys. Res. Ocean. 123, 2475–2493 (2018).ADS 

    Google Scholar 
    70.Ward, B. Near-surface ocean temperature. J. Geophys. Res. 111, C02004 (2006).ADS 

    Google Scholar 
    71.Kilpatrick, K. A., Podestá, G. P. & Evans, R. Overview of the NOAA/NASA advanced very high resolution radiometer Pathfinder algorithm for sea surface temperature and associated matchup database. J. Geophys. Res. Ocean. 106, 9179–9197 (2001).ADS 

    Google Scholar 
    72.Hollstein, A., Segl, K., Guanter, L., Brell, M. & Enesco, M. Ready-to-use methods for the detection of clouds, cirrus, snow, shadow, water and clear sky pixels in Sentinel-2 MSI images. Remote Sens. 8, 666 (2016).ADS 

    Google Scholar 
    73.Luo, B., Minnett, P. J., Gentemann, C. & Szczodrak, G. Improving satellite retrieved night-time infrared sea surface temperatures in aerosol contaminated regions. Remote Sens. Environ. 223, 8–20 (2019).ADS 

    Google Scholar 
    74.Moore, T. S., Campbell, J. W. & Dowell, M. D. A class-based approach to characterizing and mapping the uncertainty of the MODIS ocean chlorophyll product. Remote Sens. Environ. 113, 2424–2430 (2009).ADS 

    Google Scholar 
    75.Pieri, M. et al. Assessment of three algorithms for the operational estimation of [CHL] from MODIS data in the Western Mediterranean Sea. Eur. J. Remote Sens. 48, 383–401 (2015).
    Google Scholar 
    76.Tilstone, G. H. et al. Performance of Ocean Colour Chlorophyll-a algorithms for Sentinel-3 OLCI, MODIS-Aqua and Suomi-VIIRS in open-ocean waters of the Atlantic. Remote Sens. Environ. 260, 112444 (2021).ADS 

    Google Scholar  More

  • in

    First tracking of the oceanic spawning migrations of Australasian short-finned eels (Anguilla australis)

    1.Dudgeon, D. et al. Freshwater biodiversity: Importance, threats, status and conservation challenges. Biol. Rev. Camb. Philos. Soc. 81, 163–182 (2006).PubMed 

    Google Scholar 
    2.Arthington, A. H., Dulvy, N. K., Gladstone, W. & Winfield, I. J. Fish conservation in freshwater and marine realms: Status, threats and management. Aquat. Conserv. 26, 838–857 (2016).
    Google Scholar 
    3.Deinet, S. et al. The Living Planet Index (LPI) for Migratory Freshwater Fish—Technical Report. (World Fish Migration Foundation, 2020).4.Limburg, K. E. & Waldman, J. R. Dramatic declines in North Atlantic diadromous fishes. Bioscience 59, 955–965 (2009).
    Google Scholar 
    5.Lennox, R. J. et al. One hundred pressing questions on the future of global fish migration science, conservation, and policy. Front. Ecol. Evol. 7, 286 (2019).ADS 

    Google Scholar 
    6.Jellyman, D.J. An enigma: how can freshwater eels (Anguilla spp.) be such a successful genus yet be universally
    threatened? Rev. Fish Biol. Fish. https://doi.org/10.1007/s11160-021-09658-8 (2021). 7.Gross, M. R., Coleman, R. M. & McDowall, R. M. Aquatic productivity and the evolution of diadromous fish migration. Science 239, 1291–1293 (1988).ADS 
    CAS 
    PubMed 

    Google Scholar 
    8.Aarestrup, K. et al. Oceanic spawning migration of the European eel (Anguilla anguilla). Science 325, 1660–1660 (2009).ADS 
    CAS 
    PubMed 

    Google Scholar 
    9.Righton, D. et al. Empirical observations of the spawning migration of European eels: The long and dangerous road to the Sargasso Sea. Sci. Adv. 2, e1501694 (2016).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    10.Chow, S. et al. Light-sensitive vertical migration of the Japanese eel Anguilla japonica revealed by real-time tracking and its utilization for geolocation. PLoS ONE 10, e0121801 (2015).PubMed 
    PubMed Central 

    Google Scholar 
    11.Béguer-Pon, M. et al. Tracking anguillid eels: Five decades of telemetry-based research. Mar. Freshw. Res. 69, 199–219 (2018).
    Google Scholar 
    12.Jellyman, D. & Tsukamoto, K. First use of archival transmitters to track migrating freshwater eels Anguilla dieffenbachii at sea. Mar. Ecol. Prog. Ser. 233, 207–215 (2002).ADS 

    Google Scholar 
    13.Watanabe, S. et al. Reexamination of the spawning migration of Anguilla dieffenbachii in relation to water temperature and the lunar cycle. N. Z. J. Mar. Freshw. Res. 54, 131–147 (2020).
    Google Scholar 
    14.McNiven, I. et al. Phased redevelopment of an ancient Gunditjmara fish trap over the past 800 years: Muldoons Trap Complex, Lake Condah, southwestern Victoria. Aust. Archaeol. 81, 44–58 (2015).
    Google Scholar 
    15.Rose, D., Bell, D. & Crook, D. A. Restoring habitat and cultural practice in Australia’s oldest and largest traditional aquaculture system. Rev. Fish Biol. Fish. 26, 589–600 (2016).
    Google Scholar 
    16.Pike, C., Crook, V. & Gollock, M. Anguilla australis (errata version published in 2019). The IUCN Red List of Threatened Species 2019: e.T195502A154801652 (2019). https://doi.org/10.2305/IUCN.UK.2019-2.RLTS.T195502A154801652.en. Downloaded on 14 January 2020.17.Miller, M. J. et al. Review of Ocean-Atmospheric Factors in the Atlantic and Pacific Oceans Influencing Spawning and Recruitment of Anguillid Eels. 231–249 (American Fisheries Society Symposium, 2009).18.Jacoby, D. M. P. et al. Synergistic patterns of threat and the challenges facing global anguillid eel conservation. Glob. Ecol. Conserv. 4, 321–333 (2015).
    Google Scholar 
    19.Schmidt, J. The freshwater eels of Australia with some remarks on the shortfin species of Anguilla. Rec. Aust. Mus. 16, 179–210 (1928).
    Google Scholar 
    20.Jespersen, P. Indo-Pacific leptocephaids of the genus Anguilla. Systematic and biological studies. Dana-Rep. Carlsberg Found. 22, 1–128 (1942).
    Google Scholar 
    21.Castle, P. H. J. Anguillid leptocephali in the southwest Pacific. Zool. Pubs Vic. Univ. Wellingt. 33, 1–14 (1963).
    Google Scholar 
    22.Aoyama, J. et al. Distribution and dispersal of anguillid leptocephali in the western Pacific Ocean revealed by molecular analysis. Mar. Ecol. Prog. Ser. 188, 193–200 (1999).ADS 

    Google Scholar 
    23.Kuroki, M. et al. Distribution of anguillid leptocephali and possible spawning areas in the South Pacific Ocean. Progr. Oceanogr. 180, 102234 (2020).
    Google Scholar 
    24.Todd, P. R. Size and age of migrating New Zealand freshwater eels (Anguilla spp.). N. Z. J. Mar. Freshw. Res. 14, 283–293 (1980).
    Google Scholar 
    25.Sloane, R. Preliminary observations of migrating adult freshwater eels (Anguilla australis australis Richardson) in Tasmania. Mar. Freshw. Res. 35, 471–476 (1984).ADS 

    Google Scholar 
    26.Økland, F., Thorstad, E. B., Westerberg, H., Aarestrup, K. & Metcalfe, J. D. Development and testing of attachment methods for pop-up satellite archival transmitters in European eel. Anim. Biotelemetry 1, 1–13 (2013).
    Google Scholar 
    27.Kuroki, M. et al. Distribution and early life-history characteristics of anguillid leptocephali in the western South Pacific. Mar. Freshw. Res. 59, 1035–1047 (2008).
    Google Scholar 
    28.Righton, D. et al. The Anguilla spp. migration problem: 40 million years of evolution and two millennia of speculation. J. Fish Biol. 81, 365–386 (2012).CAS 
    PubMed 

    Google Scholar 
    29.Westerberg, H. Marine migratory behavior of the European silver eel. In Physiology and Ecology of Fish Migration (eds H. Ueda, H. & Tsukamoto, K.) 80–103 (CRC Press, 2013).30.Chang, Y.-L.K., Olmo, G. D. & Schabetsberger, R. Tracking the marine migration routes of South Pacific silver eels. Mar. Ecol. Prog. Ser. 646, 1–12 (2020).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    31.Westerberg, H., Sjöberg, N., Lagenfelt, I., Aarestrup, K. & Righton, D. Behaviour of stocked and naturally recruited European eels during migration. Mar. Ecol. Prog. Ser. 496, 145–157 (2014).ADS 

    Google Scholar 
    32.Ridgway, K. & Godfrey, J. Seasonal cycle of the East Australian current. J. Geophys. Res. Oceans 102, 22921–22936 (1997).ADS 

    Google Scholar 
    33.Ridgway, K. & Dunn, J. Mesoscale structure of the mean East Australian Current System and its relationship with topography. Prog. Oceanogr. 56, 189–222 (2003).ADS 

    Google Scholar 
    34.Westin, L. Migration failure in stocked eels Anguilla anguilla. Mar. Ecol. Prog. Ser. 254, 307–311 (2003).ADS 

    Google Scholar 
    35.Nordeng, H. A pheromone hypothesis for homeward migration in anadromous salmonids. Oikos 28, 155–159 (1977).CAS 

    Google Scholar 
    36.Hays, G. C., Cerritelli, G., Esteban, N., Rattray, A. & Luschi, P. Open ocean reorientation and challenges of island finding by sea turtles during long-distance migration. Curr. Biol. 30, 3236-3242 e3233 (2020).CAS 
    PubMed 

    Google Scholar 
    37.Béguer-Pon, M. et al. Shark predation on migrating adult American eels (Anguilla rostrata) in the Gulf of St. Lawrence. PLoS One 7, e46830 (2012).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    38.Wahlberg, M. et al. Evidence of marine mammal predation of the European eel (Anguilla anguilla L.) on its marine migration. Deep Sea Res. A 86, 32–38 (2014).
    Google Scholar 
    39.Béguer-Pon, M. et al. Large-scale migration patterns of silver American eels from the St. Lawrence River to the Gulf of St. Lawrence using acoustic telemetry. Can. J. Fish. Aquat. Sci. 71, 1579–1592 (2014).
    Google Scholar 
    40.Strøm, J. F. et al. Ocean predation and mortality of adult Atlantic salmon. Sci. Rep. 9, 1–11 (2019).ADS 

    Google Scholar 
    41.Hays, G. C. Tracking animals to their death. J. Anim. Ecol. 83, 5–6 (2014).PubMed 

    Google Scholar 
    42.Amilhat, E. et al. First evidence of European eels exiting the Mediterranean Sea during their spawning migration. Sci. Rep. 6, 21817 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    43.Schabetsberger, R. et al. Oceanic migration behaviour of tropical Pacific eels from Vanuatu. Mar. Ecol. Prog. Ser. 475, 177–190 (2013).ADS 

    Google Scholar 
    44.Schabetsberger, R. et al. Oceanic migration behaviour of Pacific eels from Samoa. Fish. Manag. Ecol. 26, 53–56 (2018).
    Google Scholar 
    45.Béguer-Pon, M., Shan, S., Castonguay, M. & Dodson, J. J. Behavioural variability in the vertical and horizontal oceanic migrations of silver American eels. Mar. Ecol. Prog. Ser. 585, 123–142 (2017).ADS 

    Google Scholar 
    46.Wu, K. et al. Illumination-dependent diel-vertical migration behavior in the genus Anguilla. J. Fish. Soc. Taiwan 45, 225–232 (2018).
    Google Scholar 
    47.Tesch, F. & Rohlf, N. Migration from continental waters to the spawning grounds. In Eel Biology (eds. Aida, K., Tsukamoto, K., Yamauchi, K.) 223–234. (Springer, 2003).48.Sébert, P., Scaion, D. & Belhomme, M. High hydrostatic pressure improves the swimming efficiency of European migrating silver eel. Respir. Physiol. Neurobiol. 165, 112–114 (2009).PubMed 

    Google Scholar 
    49.Jellyman, D. & Tsukamoto, K. Vertical migrations may control maturation in migrating female Anguilla dieffenbachii. Mar. Ecol. Prog. Ser. 404, 241–247 (2010).ADS 

    Google Scholar 
    50.Benoit-Bird, K. J., Dahood, A. D. & Würsig, B. Using active acoustics to compare lunar effects on predator–prey behavior in two marine mammal species. Mar. Ecol. Prog. Ser. 395, 119–135 (2009).ADS 

    Google Scholar 
    51.Owen, K., Andrews, R. D., Baird, R. W., Schorr, G. S. & Webster, D. L. Lunar cycles influence the diving behavior and habitat use of short-finned pilot whales around the main Hawaiian Islands. Mar. Ecol. Prog. Ser. 629, 193–206 (2019).ADS 

    Google Scholar 
    52.Crook, D. A. et al. Environmental cues and extended estuarine residence in seaward migrating eels (Anguilla australis). Freshw. Biol. 59, 1710–1720 (2014).
    Google Scholar 
    53.Musyl, M. K. et al. Performance of pop-up satellite archival tags. Mar. Ecol. Prog. Ser. 433, 1–28 (2011).ADS 

    Google Scholar 
    54.Weng, K. C. et al. Migration and habitat of white sharks (Carcharodon carcharias) in the eastern Pacific Ocean. Mar. Biol. 152, 877–894 (2007).
    Google Scholar 
    55.Gill, A., Bartlett, M. & Thomsen, F. Potential interactions between diadromous fishes of UK conservation importance and the electromagnetic fields and subsea noise from marine renewable energy developments. J. Fish Biol. 81, 664–695 (2012).CAS 
    PubMed 

    Google Scholar 
    56.Aarestrup, K. et al. Survival and progression rates of large European silver eel Anguilla anguilla in late freshwater and early marine phases. Aquat. Biol. 9, 263–270 (2010).
    Google Scholar 
    57.Hays, G. C. et al. Translating marine animal tracking data into conservation policy and management. Trends Ecol. Evol. 34, 459–473 (2019).PubMed 

    Google Scholar 
    58.Westerberg, H. & Wickström, H. Stock assessment of eels in the Baltic: Reconciling survey estimates to achieve quantitative analysis. ICES J. Mar. Sci. 73, 75–83 (2016).
    Google Scholar 
    59.Kaifu, K. Challenges in assessments of Japanese eel stock. Mar. Policy 102, 1–4 (2019).
    Google Scholar  More

  • in

    Urbanization favors the proliferation of Aedes aegypti and Culex quinquefasciatus in urban areas of Miami-Dade County, Florida

    1.World Health Organization. Vector-borne diseases. Available at: https://www.who.int/news-room/fact-sheets/detail/dengue-and-severe-dengue (2020).2.Wilke, A. B. B., Beier, J. C. & Benelli, G. Complexity of the relationship between global warming and urbanization—an obscure future for predicting increases in vector-borne infectious diseases. Curr. Opin. Insect Sci. 35, 1–9 (2019).PubMed 

    Google Scholar 
    3.Wilke, A. B. B. et al. Proliferation of Aedes aegypti in urban environments mediated by the availability of key aquatic habitats. Sci. Rep. 10, 12925 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    4.Wilke, A. B. B., Wilk-da-Silva, R. & Marrelli, M. T. Microgeographic population structuring of Aedes aegypti (Diptera: Culicidae). PLoS ONE 12, e0185150 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    5.Gubler, D. J. Dengue, urbanization and globalization: The unholy trinity of the 21st Century. Trop. Med. Health 39, S3–S11 (2011).
    Google Scholar 
    6.Johnson, M. T. J. & Munshi-South, J. Evolution of life in urban environments. Science 358, 8327 (2017).
    Google Scholar 
    7.Zohdy, S., Schwartz, T. S. & Oaks, J. R. The coevolution effect as a driver of spillover. Trends Parasitol. 35, 399–408 (2019).PubMed 

    Google Scholar 
    8.Rochlin, I., Faraji, A., Ninivaggi, D. V., Barker, C. M. & Kilpatrick, A. M. Anthropogenic impacts on mosquito populations in North America over the past century. Nat. Commun. 7, 13604 (2016).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    9.Wilke, A. B. B. et al. Community composition and year-round abundance of vector species of mosquitoes make Miami-Dade County, Florida a receptive gateway for arbovirus entry to the United States. Sci. Rep. 9, 8732 (2019).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    10.Burkett-Cadena, N. D. & Vittor, A. Y. Deforestation and vector-borne disease: Forest conversion favors important mosquito vectors of human pathogens. Basic Appl. Ecol. 26, 101–110 (2018).PubMed 

    Google Scholar 
    11.Rochlin, I., Harding, K., Ginsberg, H. S. & Campbell, S. R. Comparative analysis of distribution and abundance of West Nile and eastern equine encephalomyelitis virus vectors in Suffolk County, New York, using human population density and land use/cover data. J. Med. Entomol. 45, 563–571 (2008).CAS 
    PubMed 

    Google Scholar 
    12.Monaghan, A. J. et al. Consensus and uncertainty in the geographic range of Aedes aegypti and Aedes albopictus in the contiguous United States: Multi-model assessment and synthesis. PLoS Comput. Biol. 15, 1–19 (2019).
    Google Scholar 
    13.Wilke, A. B. B., Benelli, G. & Beier, J. C. Beyond frontiers: On invasive alien mosquito species in America and Europe. PLoS Negl. Trop. Dis. 14, e0007864 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    14.Kraemer, M. U. G. et al. The global compendium of Aedes aegypti and Ae. albopictus occurrence. Sci. Data 2, 150035 (2015).PubMed 
    PubMed Central 

    Google Scholar 
    15.Dirzo, R. et al. Defaunation in the anthropocene. Science 345, 401–406 (2014).ADS 
    CAS 
    PubMed 

    Google Scholar 
    16.Lewis, S. L. & Maslin, M. A. Defining the anthropocene. Nature 519, 171–180 (2015).ADS 
    CAS 
    PubMed 

    Google Scholar 
    17.Law, K. L. & Thompson, R. C. Microplastics in the seas. Science 345, 144–145 (2014).ADS 
    CAS 
    PubMed 

    Google Scholar 
    18.Jambeck, J. R. et al. Plastic waste inputs from land into the ocean. Science 347, 768–771 (2015).ADS 
    CAS 
    PubMed 

    Google Scholar 
    19.Turner, W. R., Oppenheimer, M. & Wilcove, D. S. A force to fight global warming. Nature 462, 278–279 (2009).ADS 
    CAS 
    PubMed 

    Google Scholar 
    20.United Nations. World population prospects 2019. Department of Economic and Social Affairs. World Population Prospects 2019. (2019).21.Multini, L. C., de Souza, A. L. & da S., Marrelli, M. T. & Wilke, A. B. B.,. The influence of anthropogenic habitat fragmentation on the genetic structure and diversity of the malaria vector Anopheles cruzii (Diptera: Culicidae). Sci. Rep. 10, 18018 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    22.Wilke, A. B. B. et al. Urbanization creates diverse aquatic habitats for immature mosquitoes in urban areas. Sci. Rep. 9, 15335 (2019).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    23.Pernat, N., Kampen, H., Jeschke, J. M. & Werner, D. Buzzing homes: Using citizen science data to explore the effects of urbanization on indoor mosquito communities. Insects 12, 1–13 (2021).
    Google Scholar 
    24.Blosser, E. M. & Burkett-cadena, N. D. Acta Tropica Culex (Melanoconion) panocossa from peninsular Florida, USA. Acta Trop. 167, 59–63 (2017).PubMed 

    Google Scholar 
    25.Bhatt, S. et al. The global distribution and burden of dengue. Nature 496, 504–507 (2013).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    26.Sun, K. et al. Quantifying the risk of local Zika virus transmission in the contiguous US during the 2015–2016 ZIKV epidemic. BMC Med. 16, 195 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    27.Rose, N. H. et al. Climate and urbanization drive mosquito preference for humans. Curr. Biol. 30, 3570-3579.e6 (2020).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    28.Wilke, A. B. B. et al. Mosquito adaptation to the extreme habitats of urban construction sites. Trends Parasitol. 35, 607–614 (2019).PubMed 

    Google Scholar 
    29.Ajelli, M. et al. Host outdoor exposure variability affects the transmission and spread of Zika virus: Insights for epidemic control. PLoS Negl. Trop. Dis. 11, e0005851 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    30.Mutebi, J.-P. et al. Zika virus MB16-23 in mosquitoes, Miami-Dade County, Florida, USA, 2016. Emerg. Infect. Dis. 24, 808–810 (2018).PubMed Central 

    Google Scholar 
    31.Little, E. et al. Socio-ecological mechanisms supporting high densities of Aedes albopictus (Diptera: Culicidae) in Baltimore, MD. J. Med. Entomol. 54, 1183–1192 (2017).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    32.Burkett-Cadena, N. D., McClure, C. J. W., Estep, L. K. & Eubanks, M. D. What drives the spatial distribution of mosquitoes?. Ecosphere 4, 1–16 (2013).
    Google Scholar 
    33.LaDeau, S. L., Leisnham, P. T., Biehler, D. & Bodner, D. Higher mosquito production in low-income neighborhoods of Baltimore and Washington, DC: Understanding ecological drivers and mosquito-borne disease risk in temperate cities. Int. J. Environ. Res. Public Health 10, 1505–1526 (2013).PubMed 
    PubMed Central 

    Google Scholar 
    34.Dowling, Z. et al. Linking mosquito infestation to resident socioeconomic status, knowledge, and source reduction practices in Suburban Washington, DC. EcoHealth 10, 36–47 (2013).PubMed 

    Google Scholar 
    35.Scavo, N. A., Barrera, R., Reyes-Torres, L. J. & Yee, D. A. Lower socioeconomic status neighborhoods in Puerto Rico have more diverse mosquito communities and higher Aedes aegypti abundance. J. Urban Ecol. 7, 1–11 (2021).
    Google Scholar 
    36.Trewin, B. J. et al. The elimination of the dengue vector, Aedes aegypti, from Brisbane, Australia: The role of surveillance, larval habitat removal and policy. PLoS Negl. Trop. Dis. 11, e0005848 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    37.Multini, L. C., de Souza, A. L. & da S., Marrelli, M. T. & Wilke, A. B. B.,. Population structuring of the invasive mosquito Aedes albopictus (Diptera: Culicidae) on a microgeographic scale. PLoS ONE 14, e0220773 (2019).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    38.Leta, S. et al. Global risk mapping for major diseases transmitted by Aedes aegypti and Aedes albopictus. Int. J. Infect. Dis. 67, 25–35 (2018).PubMed 

    Google Scholar 
    39.Benelli, G., Wilke, A. B. B. & Beier, J. C. Aedes albopictus (Asian Tiger Mosquito). Trends Parasitol. 36, 942–943 (2020).PubMed 

    Google Scholar 
    40.Benelli, G. & Mehlhorn, H. Declining malaria, rising of dengue and Zika virus: Insights for mosquito vector control. Parasitol. Res. 115, 1747–1754 (2016).PubMed 

    Google Scholar 
    41.Danauskas, J. X., Ehrenkranz, N. J., Davies, J. E. & Pond, W. L. Arboviruses and human disease in South Florida. Am. J. Trop. Med. Hyg. 15, 205–210 (1966).PubMed 

    Google Scholar 
    42.Gill, J., Stark, L. M. & Clark, G. G. Dengue surveillance in Florida, 1997–98. Emerg. Infect. Dis. 6, 30–35 (2000).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    43.Rey, J. Dengue in Florida (USA). Insects 5, 991–1000 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    44.Vitek, C. J., Richards, S. L., Mores, C. N., Day, J. F. & Lord, C. C. Arbovirus transmission by Culex nigripalpus in Florida, 2005. J. Med. Entomol. 45, 483–493 (2008).CAS 
    PubMed 

    Google Scholar 
    45.Messenger, A. M. et al. Serological evidence of ongoing transmission of dengue virus in permanent residents of Key West, Florida. Vector Borne Zoonotic Dis. 14, 783–787 (2014).PubMed 

    Google Scholar 
    46.Patterson, K. D. Yellow fever epidemics and mortality in the United States, 1693–1905. Soc. Sci. Med. 34, 855–865 (1992).CAS 
    PubMed 

    Google Scholar 
    47.Grubaugh, N. D. et al. Genomic epidemiology reveals multiple introductions of Zika virus into the United States. Nature 546, 401–405 (2017).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    48.Likos, A. et al. Local mosquito-borne transmission of zika virus—Miami-Dade and Broward Counties, Florida, June–August 2016. Morb. Mortal. Wkly. Rep. 65, 1032–1038 (2016).
    Google Scholar 
    49.Florida Department of Health. Available at: http://www.floridahealth.gov/diseases-and-conditions/mosquito-borne-diseases/_documents/week52arbovirusreport-12-31-16.pdf (2016).50.Florida Department of Health. Available at: http://www.floridahealth.gov/diseases-and-conditions/mosquito-borne-diseases/_documents/alert-dade-wnv-human-10-19-20.pdf (2020)51.Wilke, A. B. B. et al. Local conditions favor dengue transmission in the contiguous United States. Entomol. Gen. 41, 523–529 (2021).
    Google Scholar 
    52.Alto, B. W., Connelly, C. R., O’Meara, G. F., Hickman, D. & Karr, N. Reproductive biology and susceptibility of Florida Culex coronator to infection with West Nile virus. Vector-Borne Zoonotic Dis. 14, 606–614 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    53.Honório, N. A., Wiggins, K., Câmara, D. C. P., Eastmond, B. & Alto, B. W. Chikungunya virus vector competency of Brazilian and Florida mosquito vectors. PLoS Negl. Trop. Dis. 12, 1–16 (2018).
    Google Scholar 
    54.Richards, S. L., Anderson, S. L. & Lord, C. C. Vector competence of Culex pipiens quinquefasciatus (Diptera: Culicidae) for West Nile virus isolates from Florida. Trop. Med. Int. Heal. 19, 610–617 (2014).
    Google Scholar 
    55.Hribar, L. J., Smith, J. M., Vlach, J. J. & Verna, T. N. Survey of container-breeding mosquitoes from the Florida Keys, Monroe County, Florida. J. Am. Mosq. Control Assoc. 17, 245–248 (2001).CAS 
    PubMed 

    Google Scholar 
    56.United States Environmental Protection Agency. Growing for a sustainable future: Miami-Dade County urban development boundary assessment. Available at: http://www.epa.gov/smartgrowth/pdf/Miami-Dade_Final_Report_12-12-12.pdf (2012).57.Miami-Dade County Building Permits. Available at, http://www.miamidade.gov/permits/.58.Wilke, A. B. B., Carvajal, A., Vasquez, C., Petrie, W. D. & Beier, J. C. Urban farms in Miami-Dade County, Florida have favorable environments for vector mosquitoes. PLoS ONE 15, e0230825 (2020).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    59.Reba, M., Reitsma, F. & Seto, K. C. Spatializing 6,000 years of global urbanization from 3700 BC to AD 2000. Sci. Data 3, 1–16 (2016).
    Google Scholar 
    60.Ceretti-Júnior, W. et al. Mosquito faunal survey in a central park of the city of São Paulo, Brazil. J. Am. Mosq. Control Assoc. 31, 172–176 (2015).PubMed 

    Google Scholar 
    61.Ferraguti, M. et al. Effects of landscape anthropization on mosquito community composition and abundance. Sci. Rep. 6, 29002 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    62.Zahouli, J. B. Z. et al. Effect of land-use changes on the abundance, distribution, and host-seeking behavior of Aedes arbovirus vectors in oil palm-dominated landscapes, southeastern Côte d’Ivoire. PLoS ONE 12, e0189082 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    63.Westby, K. M., Adalsteinsson, S. A., Biro, E. G., Beckermann, A. J. & Medley, K. A. Aedes albopictus populations and larval habitat characteristics across the landscape: Significant differences exist between urban and rural land use types. Insects 12, 196 (2021).PubMed 
    PubMed Central 

    Google Scholar 
    64.Estallo, E. L. et al. Modelling the distribution of the vector Aedes aegypti in a central Argentine city. Med. Vet. Entomol. 32, 451–461 (2018).CAS 
    PubMed 

    Google Scholar 
    65.Messina, J. P. et al. A global compendium of human dengue virus occurrence. Sci. Data 1, 140004 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    66.Cunha, M. S. et al. Epizootics due to yellow fever virus in São Paulo State, Brazil: viral dissemination to new areas (2016–2017). Sci. Rep. 9, 5474 (2019).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    67.Ronca, S. E., Murray, K. O. & Nolan, M. S. Cumulative incidence of West Nile virus infection, continental United States, 1999–2016. Emerg. Infect. Dis. 25, 325–327 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    68.Poletti, P. et al. Transmission potential of chikungunya virus and control measures: The case of Italy. PLoS ONE 6, e18860 (2011).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    69.Wilk-da-Silva, R. & de Souza Leal Diniz, M. M. C., Marrelli, M. T. & Wilke, A. B. B.,. Wing morphometric variability in Aedes aegypti (Diptera: Culicidae) from different urban built environments. Parasit. Vectors 11, 561 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    70.Wilke, A. B. B. et al. Cemeteries in Miami-Dade County, Florida are important areas to be targeted in mosquito management and control efforts. PLoS ONE 15, e0230748 (2020).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    71.Weaver, S. C. Urbanization and geographic expansion of zoonotic arboviral diseases: Mechanisms and potential strategies for prevention. Trends Microbiol. 21, 360–363 (2013).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    72.Wilke, A. B. B., Vasquez, C., Petrie, W. & Beier, J. C. Tire shops in Miami-Dade County, Florida are important producers of vector mosquitoes. PLoS ONE 14, 2 (2019).
    Google Scholar 
    73.Kothera, L., Godsey, M., Mutebi, J. P. & Savage, H. M. A comparison of aboveground and belowground populations of Culex pipiens (Diptera: Culicidae) mosquitoes in Chicago, Illinois, and New York City, New York, using microsatellites. J. Med. Entomol. 47, 805–813 (2010).PubMed 

    Google Scholar 
    74.World Health Organization. Handbook for Integrated Vector Management (World Health Organization, 2012).
    Google Scholar 
    75.Lizzi, K. M., Qualls, W. A., Brown, S. C. & Beier, J. C. Expanding Integrated Vector Management to promote healthy environments. Trends Parasitol. 30, 394–400 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    76.Souza, R. L. et al. Effect of an intervention in storm drains to prevent Aedes aegypti reproduction in Salvador, Brazil. Parasit. Vectors 10, 1–6 (2017).
    Google Scholar 
    77.Wilke, A. B. B., Beier, J. C. & Benelli, G. Transgenic mosquitoes—Fact or fiction?. Trends Parasitol. 34, 456–465 (2018).PubMed 

    Google Scholar 
    78.Beier, J. C., Wilke, A. B. B. & Benelli, G. Newer approaches for malaria vector control and challenges of outdoor transmission. Towards Malaria Elimination – A Leap Forward https://doi.org/10.5772/intechopen.75513 (2018).Article 

    Google Scholar 
    79.World Health Organization. Tenth Meeting of the WHO Vector Control Advisory Group. (2019).80.Wilke, A. B. B. et al. Effectiveness of adulticide and larvicide in controlling high densities of Aedes aegypti in urban environments. PLoS ONE 16, e0246046 (2021).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    81.Vicente-Serrano, S. M. et al. Response of vegetation to drought time-scales across global land biomes. Proc. Natl. Acad. Sci. 110, 52–57 (2013).ADS 
    CAS 
    PubMed 

    Google Scholar 
    82.Rifat, S. A. & Al & Liu, W.,. Quantifying spatiotemporal patterns and major explanatory factors of urban expansion in Miami metropolitan area during 1992–2016. Remote Sens. 11, 2493 (2019).ADS 

    Google Scholar 
    83.Fuller, D. O. & Wang, Y. Recent trends in satellite vegetation index observations indicate decreasing vegetation biomass in the southeastern saline Everglades wetlands. Wetlands 34, 67–77 (2014).
    Google Scholar 
    84.Wilke, A. B. B. et al. Assessment of the effectiveness of BG-Sentinel traps baited with CO2 and BG-Lure for the surveillance of vector mosquitoes in Miami-Dade County. Florida. PLoS One 14, e0212688 (2019).CAS 
    PubMed 

    Google Scholar 
    85.Darsie, R. F. Jr. & Morris, C. D. Keys to the adult females and fourth-instar larvae of the mosquitoes of Florida (Diptera, Culicidae). 1st ed. Vol. 1. Tech Bull Florida Mosq Cont Assoc (2000).86.Anderson, M. J. Permutational Multivariate Analysis of Variance (PERMANOVA). Wiley StatsRef: Statistics Reference Online. 1–15 (2017) DOI:https://doi.org/10.1002/9781118445112.stat07841.87.Alencar, J. et al. Culicidae community composition and temporal dynamics in Guapiaçu ecological reserve, Cachoeiras de Macacu, Rio de Janeiro, Brazil. PLoS ONE 10, 1–16 (2015).
    Google Scholar 
    88.Clarke, K. R. Non-parametric multivariate analyses of changes in community structure. Austral Ecol. 18, 117–143 (1993).
    Google Scholar 
    89.Hammer, Ø., Harper, D. A. T. T. & Ryan, P. D. PAST: Paleontological statistics software package for education and data analysis. Palaeontol. Electron. 4, 9 (2001).
    Google Scholar 
    90.Ryan, P. A., Lyons, S. A., Alsemgeest, D., Thomas, P. & Kay, B. H. Spatial statistical analysis of adult mosquito (Diptera: Culicidae) counts: An example using light trap data, in Redland Shire, southeastern Queensland, Australia. J. Med. Entomol. 41, 1143–1156 (2004).PubMed 

    Google Scholar 
    91.O’Brien, R. M. A caution regarding rules of thumb for variance inflation factors. Qual. Quant. 41, 673–690 (2007).
    Google Scholar 
    92.Wilke, A. B. B., Medeiros-Sousa, A. R., Ceretti-Junior, W. & Marrelli, M. T. Mosquito populations dynamics associated with climate variations. Acta Trop. 166, 343–350 (2016).PubMed 

    Google Scholar 
    93.Cohen, J. Eta-squared and partial eta-squared in fixed factor ANOVA designs. Educ. Psychol. Meas. 33, 107–112 (1973).
    Google Scholar  More

  • in

    Trajectory of body mass index and height changes from childhood to adolescence: a nationwide birth cohort in Japan

    ParticipantsThe Ministry of Health, Labour, and Welfare of Japan has been conducting The Longitudinal Survey of Newborns in the 21st Century since 2001 to establish strategies to counter the declining birthrate in Japan. The survey targeted all babies born in Japan between January 10 and 17 or between July 10 and 17 of 2001. Baseline questionnaires were sent to a total of 53,575 families when eligible babies reached the age of 6 months and 47,015 families initially completed the baseline questionnaire (88% response rate). These respondents were mailed follow-up questionnaires to investigate medical conditions and behaviors when children reached the ages of 1.5, 2.5, 3.5, 4.5, 5.5, 7, 8, 9, 10, 11, 12, 13, 14, and 15 years20,21,22,23. Birth record data from Vital Statistics of Japan are also linked for each child participating in the study. The current study included data for children/families who responded both to the baseline questionnaire and the fifteenth questionnaire at age 15 years.The baseline survey at age 6 months included questions regarding children’s perinatal status as well as household and socioeconomic factors such as parental academic attainment, parental smoking status, and daycare attendance. The subsequent annual surveys starting at age 1.5 years included questions regarding each child’s height, weight and health status. We excluded 2382 children born before 37 weeks of pregnancy and one child with responses only for the baseline survey and the survey at age 15 years. A total of 26,778 children (315,581 data points) were included in the final analysis. A total of 11,141 children (41.61%) had responses to all 15 questionnaires between the ages of 6 months and 15 years, and responses to more than 12 questionnaires were available for the majority (91.94%) of children (Fig. 1, Table S1).Figure 1Flowchart of study participants.Full size imageMeasuresWe calculated BMI based on each participant’s reported annual height and weight. Each participant’s annual BMI was converted to a BMI Z-score using smoothed L, M, and S values for BMI standards from a representative population of Japanese children24. Briefly, the LMS (lambda–mu–sigma) method is a method proposed by Cole et al. to monitor changes in the skewness of the distribution during childhood as a way of constructing normalized growth standards25. Participants were then classified into four BMI categories based on the World Health Organization (WHO) criteria26: underweight (BMI standard deviation [SD] score of − 5 or more but less than − 2), normal weight (BMI SD score of − 2 or more but less than 1), overweight (BMI SD score of 1 or more but less than 2), and obese (BMI SD score of 2 or more but less than 5). The definitions of overweight and obesity were different for children under 5 years of age: a BMI Z-score of 2 SD or more was categorized as overweight and a BMI Z-score of 3 SD or more was categorized as obese. BMI category at age 15 years was the main outcome of interest in the current study.We also calculated annual height growth for each participant by subtracting the height reported at the previous survey from that reported in the current survey. For annual height growth between 5.5 and 7 years of age, this value was multiplied by 2/3 because of the 1.5-year interval between surveys.Statistical analysesWe first compared baseline characteristics among the four BMI categories (underweight, normal weight, overweight and obese) at age 15 years. To evaluate potential selection bias resulting from losses to follow-up, we also compared the baseline characteristics of children included in the analysis and those of children lost to follow-up through to the fifteenth survey (at age 15 years).We retrospectively examined annual aggregate categorical changes in individuals of the four BMI categories (groups) at age 15 years. For each group, the proportion of each BMI category at each survey between the ages of 1.5 and 14 years was calculated. In addition, we prospectively calculated the proportion of children in each BMI category at each survey between the ages of 1.5 and 14 years who eventually became underweight, normal weight, overweight, or obese at age 15 years. Note that these analyses were based on aggregate data and do not describe individual BMI changes and were performed using only the data obtained without imputation of missing values.Under the assumption that missing data were missing at random, mixed effect models with natural cubic regression splines were applied to calculate the trajectories of BMI Z-scores and annual BMI Z-score changes through age 15 years for participants of each BMI category at age 15 years. Knots at seven locations were placed in percentiles of age to yield a sufficient number of measurements between each consecutive knot (age 1.5, 3.5, 5.5, 8.5, 11, 13 and 15 years), as recommended by Harrell27. The mixed effect model is useful for describing population average growth trajectories and individual growth trajectories even when data are not available for all children at all ages28,29,30,31. Briefly, the population average growth trajectory was modeled with fixed effects, while the individual variability is represented as random effects.After fitting individual BMI trajectories using a mixed-effects model with natural cubic spline function, we estimated individual adiposity rebound timing as the age where the first derivative of the trajectory reached its minimum and the second derivative was positive32. Children were then classified into five categories (1.5–2.5 years, 3.5–4.5 years, 5.5–7 years, 8–10 years, and 11 years or older) for analysis of adiposity rebound timing33,34. The distribution of adiposity rebound timing was calculated for individuals of each BMI status at age 15 years overall and by gender.Finally, we modelled annual height change and its associations with BMI status at age 15 years separately for each gender using mixed-effects models with natural cubic regression splines.All statistical analyses were performed using Stata version 16 (StataCorp LLC, College Station, TX, USA). This study was approved by the Institutional Review Board at Okayama University Graduate School of Medicine, Dentistry, and Pharmaceutical Sciences (No.1506-073) and was conducted in accordance with the 1964 Helsinki Declaration and Ethical Guidelines for Medical and Health Research Involving Human Subjects. Informed consent was obtained by the opt-out method on the university’s website. More

  • in

    Translation stalling proline motifs are enriched in slow-growing, thermophilic, and multicellular bacteria

    1.Russell JB, Cook GM. Energetics of bacterial growth: balance of anabolic and catabolic reactions. Microbiol Rev. 1995;59:48–62.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    2.Klumpp S, Scott M, Pedersen S, Hwa T. Molecular crowding limits translation and cell growth. PNAS. 2013;110:16754–9.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    3.Pedersen S. Escherichia coli ribosomes translate in vivo with variable rate. EMBO J. 1984;3:2895–8.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    4.Ran W, Higgs PG. Contributions of speed and accuracy to translational selection in bacteria. PLoS One. 2012;7:e51652.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    5.Vieira-Silva S, Rocha E. The systemic imprint of growth and its uses in ecological (meta)genomics. PLoS Genet. 2009;6:1–15.
    Google Scholar 
    6.Roller BRK, Stoddard SF, Schmidt TM. Exploiting rRNA operon copy number to investigate bacterial reproductive strategies. Nat Microbiol. 2016;1:1–7.
    Google Scholar 
    7.Buskirk AR, Green R. Ribosome pausing, arrest and rescue in bacteria and eukaryotes. Philos Trans R Soc B. 2017;372:20160183–11.
    Google Scholar 
    8.Wohlgemuth I, Brenner S, Beringer M, Rodnina MV. Modulation of the rate of peptidyl transfer on the ribosome by the nature of substrates. J Biol Chem. 2008;283:32229–35.CAS 
    PubMed 

    Google Scholar 
    9.Pavlov MY, Watts RE, Tan Z, Cornish VW, Ehrenberg M, Forster AC. Slow peptide bond formation by proline and other N-alkylamino acids in translation. PNAS. 2009;106:50–54.CAS 
    PubMed 

    Google Scholar 
    10.Mandal A, Mandal S, Park MH. Genome-wide analyses and functional classification of proline repeat-rich proteins: potential role of eIF5A in eukaryotic evolution. PLoS One. 2014;9:e111800–13.PubMed 
    PubMed Central 

    Google Scholar 
    11.Adzhubei AA, Sternberg MJE, Makarov AA. Polyproline-II helix in proteins: structure and function. J Mol Biol. 2013;425:2100–32.CAS 
    PubMed 

    Google Scholar 
    12.Elam WA, Schrank TP, Campagnolo AJ, Hilser VJ. Evolutionary conservation of the polyproline II conformation surrounding intrinsically disordered phosphorylation sites. Protein Sci. 2013;22:405–17.PubMed 

    Google Scholar 
    13.Ball LJ, Kühne R, Schneider-Mergener J, Oschkinat H. Recognition of proline-rich motifs by protein-protein-interaction domains. Angew Chem Int Ed Engl. 2005;44:2852–69.CAS 
    PubMed 

    Google Scholar 
    14.Starosta AL, Lassak J, Peil L, Atkinson GC, Virumäe K, Tenson T, et al. Translational stalling at polyproline stretches is modulated by the sequence context upstream of the stall site. Nucleic Acids Res. 2014;42:10711–9.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    15.Woolstenhulme CJ, Guydosh NR, Green R, Buskirk AR. High-precision analysis of translational pausing by ribosome profiling in bacteria lacking EFP. Cell Rep. 2015;11:13–21.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    16.Hersch SJ, Elgamal S, Katz A, Ibba M, Navarre WW. Translation initiation rate determines the impact of ribosome stalling on bacterial protein synthesis. J Biol Chem. 2014;289:28160–71.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    17.Lassak J, Wilson DN, Jung K. Stall no more at polyproline stretches with the translation elongation factors EF‐P and IF‐5A. Mol Microbiol. 2016;99:219–35.CAS 
    PubMed 

    Google Scholar 
    18.Yanagisawa T, Sumida T, Ishii R, Takemoto C, Yokoyama S. A paralog of lysyl-tRNA synthetase aminoacylates a conserved lysine residue in translation elongation factor P. Nature. 2010;17:1136–43.CAS 

    Google Scholar 
    19.Park J-H, Johansson HE, Aoki H, Huang BX, Kim H-Y, Ganoza MC, et al. Post-translational modification by beta-lysylation is required for activity of Escherichia coli elongation factor P (EF-P). J Biol Chem. 2012;287:2579–90.CAS 
    PubMed 

    Google Scholar 
    20.Lassak J, Keilhauer E, Fürst M, Wuichet K, Gödeke J, Starosta AL, et al. Arginine-rhamnosylation as new strategy to activate translation elongation factor P. Nat Chem Biol. 2015;11:266–70.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    21.Tollerson R, Witzky A, Ibba M. Elongation factor P is required to maintain proteome homeostasis at high growth rate. PNAS. 2018;115:1–6.
    Google Scholar 
    22.Peng WT, Banta LM, Charles TC, Nester EW. The chvH locus of Agrobacterium encodes a homologue of an elongation factor involved in protein synthesis. J Bacteriol. 2001;183:36–45.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    23.Rajkovic A, Hummels KR, Witzky A, Erickson S, Gafken PR, Whitelegge JP, et al. Translation control of swarming proficiency in Bacillus subtilis by 5-amino-pentanolylated elongation factor P. J Biol Chem. 2016;291:10976–85.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    24.Navarre WW, Zou SB, Roy H, Xie JL, Savchenko A, Singer A, et al. PoxA, YjeK, and elongation factor P coordinately modulate virulence and drug resistance in Salmonella enterica. Mol Cell. 2010;39:209–21.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    25.Hummels KR, Kearns DB. Suppressor mutations in ribosomal proteins and FliY restore Bacillus subtilis swarming motility in the absence of EF-P. PLoS Genet. 2019;15:e1008179–27.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    26.Rajkovic A, Erickson S, Witzky A, Branson OE, Seo J, Gafken PR, et al. Cyclic rhamnosylated elongation factor P establishes antibiotic resistance in Pseudomonas aeruginosa. MBio. 2015;6:1–9.
    Google Scholar 
    27.Yanagisawa T, Takahashi H, Suzuki T, Masuda A, Dohmae N, Yokoyama S. Neisseria meningitidis translation elongation factor P and its active-site arginine residue are essential for cell viability. PLoS One. 2016;11:e0147907–27.PubMed 
    PubMed Central 

    Google Scholar 
    28.Krafczyk R, Qi F, Sieber A, Mehler J, Jung K, Frishman D, et al. Proline codon pair selection determines ribosome pausing strength and translation efficiency in bacteria. Commun Biol. 2021;4:1–11.
    Google Scholar 
    29.Qi F, Motz M, Jung K, Lassak J, Frishman D. Evolutionary analysis of polyproline motifs in Escherichia coli reveals their regulatory role in translation. PLoS Comput Biol. 2018;14:e1005987–19.PubMed 
    PubMed Central 

    Google Scholar 
    30.Karlin S, Mrázek J, Campbell A, Kaiser D. Characterizations of highly expressed genes of four fast-growing bacteria. J Bacteriol. 2001;183:5025–40.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    31.Dethlefsen L, Schmidt TM. Performance of the translational apparatus varies with the ecological strategies of bacteria. J Bacteriol. 2007;189:3237–45.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    32.Weissman JL, Hou S, Fuhrman JA. Estimating maximal microbial growth rates from cultures, metagenomes, and single cells via codon usage patterns. PNAS. 2021;118:1–10.
    Google Scholar 
    33.Hersch SJ, Wang M, Zou SB, Moon K-M, Foster LJ, Ibba M, et al. Divergent protein motifs direct elongation factor P-mediated translational regulation in Salmonella enterica and Escherichia coli. MBio. 2013;4:1–10.
    Google Scholar 
    34.Pinheiro B, Scheidler CM, Kielkowski P, Schmid M, Forné I, Ye S, et al. Structure and function of an elongation factor P subfamily in Actinobacteria. Cell Rep. 2020;30:4332–42. e5.CAS 
    PubMed 

    Google Scholar 
    35.Chen I-MA, Markowitz VM, Chu K, Palaniappan K, Szeto E, Pillay M, et al. IMG/M: integrated genome and metagenome comparative data analysis system. Nucleic Acids Res. 2017;45:D507–D516.CAS 
    PubMed 

    Google Scholar 
    36.Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 2015;25:1043–55.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    37.Chaumeil P-A, Mussig AJ, Hugenholtz P, Parks DH. GTDB-Tk: a toolkit to classify genomes with the Genome Taxonomy Database. Bioinformatics. 2020;36:1925–7.CAS 

    Google Scholar 
    38.Eddy SR. Accelerated profile HMM searches. PLoS Comput Biol. 2011;7:e1002195–16.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    39.Madin JS, Nielsen DA, Brbic M, Corkrey R, Danko D, Edwards K, et al. A synthesis of bacterial and archaeal phenotypic trait data. Sci Data. 2020;7:1–8.
    Google Scholar 
    40.Novembre JA. Accounting for background nucleotide composition when measuring codon usage bias. Mol Biol Evol. 2002;19:1390–4.CAS 
    PubMed 

    Google Scholar 
    41.Erdos G, Dosztányi Z. Analyzing protein disorder with IUPred2A. Curr Protoc Bioinforma. 2020;70:1–15.
    Google Scholar 
    42.Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004;32:1792–7.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    43.Price MN, Dehal PS, Arkin AP. FastTree: computing large minimum evolution trees with profiles instead of a distance matrix. Mol Biol Evol. 2009;26:1641–50.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    44.Pagel M. Inferring the historical patterns of biological evolution. Nature. 1999;401:877–84.CAS 
    PubMed 

    Google Scholar 
    45.Revell LJ. phytools: an R package for phylogenetic comparative biology (and other things). Methods Ecol Evol. 2011;3:217–23.
    Google Scholar 
    46.Orme D, Freckleton R, Thomas G, Petzoldt T, Fritz S, Isaac N, et al. caper: comparative analyses of phylogenetics and evolution in R. 2018; https://CRAN.R-project.org/package=caper.47.Wickham H. ggplot2: elegant graphics for data analysis. 2016. Springer-Verlag New York.48.Symonds MRE, Blomberg SP. A primer on phylogenetic generalized least squares. In: Garamszegi L (eds). Modern phylogenetic comparative methods and their application in evolutionary biology. (Springer, Berlin, Heidelberg, 2014) pp. 105–30.49.Watanabe K, Suzuki Y. Protein thermostabilization by proline substitutions. J Mol Catal B Enzym. 1998;4:167–80.CAS 

    Google Scholar 
    50.Sabath N, Ferrada E, Barve A, Wagner A. Growth temperature and genome size in bacteria are negatively correlated, suggesting genomic streamlining during thermal adaptation. Genome Biol Evol. 2013;5:966–77.PubMed 
    PubMed Central 

    Google Scholar 
    51.Goldman BS, Nierman WC, Kaiser D, Slater SC, Durkin AS, Eisen JA, et al. Evolution of sensory complexity recorded in a myxobacterial genome. PNAS. 2006;103:15200–5.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    52.Long AM, Hou S, Ignacio-Espinoza JC, Fuhrman JA. Benchmarking microbial growth rate predictions from metagenomes. ISME J. 2020;15:1–13.
    Google Scholar 
    53.Rocha EPC. Codon usage bias from tRNA’s point of view: redundancy, specialization, and efficient decoding for translation optimization. Genome Res. 2004;14:2279–86.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    54.Klappenbach JA, Dunbar JM, Schmidt TM. rRNA operon copy number reflects ecological strategies of bacteria. Appl Environ Microbiol. 2000;66:1328–33.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    55.Kanehisa M, Goto S, Sato Y, Furumichi M, Tanabe M. KEGG for integration and interpretation of large-scale molecular data sets. Nucleic Acids Res. 2012;40:D109–D114.CAS 
    PubMed 

    Google Scholar 
    56.Perez J, Castaneda-García A, Jenke-Kodama H, Muller R, Munoz-Dorado J. Eukaryotic-like protein kinases in the prokaryotes and the myxobacterial kinome. PNAS. 2008;105:15950–5.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    57.Shi L, Pigeonneau N, Ravikumar V, Dobrinic P, Macek B, Franjevic D, et al. Cross-phosphorylation of bacterial serine/threonine and tyrosine protein kinases on key regulatory residues. Front Microbiol. 2014;5:1–13.CAS 

    Google Scholar 
    58.Jakob U, Kriwacki R, Uversky VN. Conditionally and transiently disordered proteins: awakening cryptic disorder to regulate protein function. Chem Rev. 2014;114:6779–805.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    59.Starosta AL, Lassak J, Peil L, Atkinson GC, Woolstenhulme CJ, Virumäe K, et al. A conserved proline triplet in Val-tRNA synthetase and the origin of elongation factor P. Cell Rep. 2014;9:476–83.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    60.Nariya H, Inouye S. A protein Ser/Thr kinase cascade negatively regulates the DNA-binding activity of MrpC, a smaller form of which may be necessary for the Myxococcus xanthus development. Mol Microbiol. 2006;60:1205–17.CAS 
    PubMed 

    Google Scholar 
    61.Stein EA, Cho K, Higgs PI, Zusman DR. Two Ser/Thr protein kinases essential for efficient aggregation and spore morphogenesis in Myxococcus xanthus. Mol Microbiol. 2006;60:1414–31.CAS 
    PubMed 

    Google Scholar 
    62.Iakoucheva LM, Radivojac P, Brown CJ, OConnor TR, Sikes JG, Obradovic Z, et al. The importance of intrinsic disorder for protein phosphorylation. Nucleic Acids Res. 2004;32:1037–49.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    63.Elsen S, Swem LR, Swem DL, Bauer CE. RegB/RegA, a highly conserved redox-responding global two-component regulatory system. Microbiol Mol Biol R. 2004;68:263–79.CAS 

    Google Scholar 
    64.Tawa P, Stewart RC. Kinetics of CheA autophosphorylation and dephosphorylation reactions. Biochemistry. 1994;33:7917–24.CAS 
    PubMed 

    Google Scholar 
    65.Yoshida T, jian CaiS, Inouye M. Interaction of EnvZ, a sensory histidine kinase, with phosphorylated OmpR, the cognate response regulator. Mol Microbiol. 2002;46:1283–94.CAS 
    PubMed 

    Google Scholar 
    66.Cho M-H, Wrabl JO, Taylor J, Hilser VJ. Hidden dynamic signatures drive substrate selectivity in the disordered phosphoproteome. PNAS. 2020;117:1–11.
    Google Scholar  More

  • in

    Rank-invariant estimation of inbreeding coefficients

    Statistical samplingWe can describe the dependence between pairs of uniting alleles in a single population without invoking an evolutionary model for the history of the population. In this “statistical sampling” framework (Weir, 1996) we do not consider the variation associated with evolutionary processes but we do consider the variation among samples from the same population. Although extensive sets of genetic data allow individual-level inbreeding coefficients to be estimated with high precision, we start with population-level estimation.Allelic dependencies can be quantified with the within-population inbreeding coefficient, written here as fW to emphasize it is a within-population quantity, defined by$${H}_{l}=2{p}_{l}(1-{p}_{l})(1-{f}_{W})$$
    (1)
    where Hl is the population proportion of heterozygotes for the reference allele at SNP l and pl is the population proportion of that allele. The same value of fW is assumed to apply for all SNPs. An immediate consequence of this definition is that the population proportions of homozygotes for the reference and alternative alleles are ({p}_{l}^{2}+{p}_{l}(1-{p}_{l}){f}_{W}) and ({(1-{p}_{l})}^{2}+{p}_{l}(1-{p}_{l}){f}_{W}) respectively. This formulation allows fW to be negative, with the maximum of −pl/(1 − pl) and −(1 − pl)/pl as lower bound. It is bounded above by 1. Hardy–Weinberg equilibrium, HWE, corresponds to fW = 0 and textbooks (e.g., (Hedrick, 2000)) point out that negative values of fW indicate more heterozygotes than expected under HWE.Observed heterozygote proportions ({tilde{H}}_{l}) have Hl as within-population expectation ({{{{{{mathcal{E}}}}}}}_{W}) over samples from the study population, ({{{{{{mathcal{E}}}}}}}_{W}({tilde{H}}_{l})={H}_{l}), and this would provide a simple estimator of fW if the population allele proportions were known. In practice, however, these proportions are unknown. Steele et al. (2014) suggested use of data external to the study sample to provide reference allele proportions in forensic applications where a reference database is used for making inferences about the population relevant for a particular crime. The more usual approach is to use study sample proportions ({tilde{p}}_{l}) in place of the true proportions pl, as in equation 1 of Li & Horvitz (1953):$${hat{f}}_{{W}_{l}}=1-frac{{tilde{H}}_{l}}{2{tilde{p}}_{l}(1-{tilde{p}}_{l})}$$
    (2)
    The moment estimator in Eq. (2) is also an MLE of fW when only one locus is considered, but it is biased (Robertson & Hill, 1984) since not only is it a ratio of statistics but also the expected value ({{{{{{mathcal{E}}}}}}}_{W}[2{tilde{p}}_{l}(1-{tilde{p}}_{l})]) over repeated samples of n from the population is 2pl(1 − pl)[1 − (1 + fW)/(2n)] (e.g., (Weir, 1996), p39).This approach can be used to estimate the within-population inbreeding coefficient fj for each individual j in a sample from one population. These are the “simple” estimators of Hall et al. (2012) and the ({hat{f}}_{{{{{{{rm{HOM}}}}}}}_{j}}) of Yengo et al. (2017):$${hat{f}}_{{{{{{{rm{HOM}}}}}}}_{jl}}=1-frac{{tilde{H}}_{jl}}{2{tilde{p}}_{l}(1-{tilde{p}}_{l})}$$
    (3)
    The sample heterozygosity indicator ({tilde{H}}_{jl}) is one if individual j is heterozygous at SNP l and is zero otherwise. Averaging Eq. (3) over individuals gives the estimator based on SNP l in Eq. (2).A single SNP provides estimates that are either 1 or a negative value depending on ({tilde{p}}_{l}), so many SNPs are used in practice. In both Hall et al. (2012) and Yengo et al. (2017) data were combined over loci as weighted or “ratio of averages” estimators:$${hat{f}}_{{{{{{{rm{Hom}}}}}}}_{j}}=1-frac{{sum }_{l}({tilde{H}}_{jl})}{{sum }_{l}[2{tilde{p}}_{l}(1-{tilde{p}}_{l})]}$$
    (4)
    Gazal et al. (2014) referred to this estimator as fPLINK as it is an option in PLINK. We show below the good performance of this weighted estimator for large sample sizes and large numbers of loci. We will consider throughout that a large number L of SNPs are used so that ratios of sums of statistics over loci, such as in Eq. (4), have expected values equal to the ratio of expected values of their numerators and denominators. Ochoa & Storey (2021) showed statistics of the form ({tilde{A}}_{L}/{tilde{B}}_{L}), where ({tilde{A}}_{L}=mathop{sum }nolimits_{l = 1}^{L}{a}_{l}/L) and ({tilde{B}}_{L}=mathop{sum }nolimits_{l = 1}^{L}{b}_{l}/L), have expected values that converge almost surely to the ratio A/B when ({{{{{{mathcal{E}}}}}}}_{W}({tilde{A}}_{L})=A{c}_{L}) and ({{{{{{mathcal{E}}}}}}}_{W}({tilde{B}}_{L})=B{c}_{L}). This result rests on the expectations ({{{{{{mathcal{E}}}}}}}_{W}({a}_{l})=A{c}_{l}) and ({{{{{{mathcal{E}}}}}}}_{W}({b}_{l})=B{c}_{l}) with ({c}_{L}=mathop{sum }nolimits_{l = 1}^{L}{c}_{l}/L). It requires ∣al∣, ∣bl∣ to both be no greater than some finite quantity C, cL to converge to a finite value c as L increases, and for Bc not to be zero. For the ratio in Eq. (4), ({a}_{l}={tilde{H}}_{jl}), ({b}_{l}=2{tilde{p}}_{l}(1-{tilde{p}}_{l})) so A = (1 − fj), B = 1 for large sample sizes n, and cL = ∑l2pl(1 − pl)/L ≤ 1/2. The conditions are satisfied providing at least one SNP is polymorphic. For an “average of ratios” estimator of the form (mathop{sum }nolimits_{l = 1}^{L}({a}_{l}/{b}_{l})/L), the denominators bl can be very small and convergence of its expected value is not assured.As an alternative to using sample allele frequencies, Hall et al. (2012) used maximum likelihood to estimate population allele proportions for multiple loci whereas Ayres & Balding (1998) used Markov chain Monte Carlo methods in a Bayesian approach that integrated out the allele proportion parameters. Neither of those papers considered data of the size we now face in sequence-based studies of many organisms, and we doubt the computational effort to estimate, or integrate over, hundreds of millions of allele proportions in Eqs. (2) or (4) adds much value to inferences about f. The allele-sharing estimators we describe below regard allele probabilities as unknown nuisance parameters and we show how to avoid estimating them or assigning them values.Hall et al. (2012) used an EM algorithm to find MLEs for fj when population allele proportions were regarded as being known and equal to sample proportions. Alternatively, a grid search can be conducted over the range of validity for the single parameter fj that maximizes the log-likelihood$${{{{mathrm{ln}}}}},[{{{{{rm{Lik}}}}}}({f}_{j})]={{{{{rm{Constant}}}}}}+mathop{sum }limits_{l=1}^{L}{{tilde{H}}_{jl}{{{{mathrm{ln}}}}},[(1-{f}_{j})]+(1-{tilde{H}}_{jl}){{{{mathrm{ln}}}}},[1-2{tilde{p}}_{l}(1-{tilde{p}}_{l})(1-{f}_{j})]}$$Estimation of the within-population inbreeding coefficients fW (FIS of (Wright, 1922)) and fj does not require any information beyond genotype proportions in samples from a study population, nor does it make any assumptions about that population or the evolutionary forces that shaped the population. The coefficients are simply measures of dependence of pairs of alleles within individuals.Genetic samplingInbreeding parameters of most interest in genetic studies are those that recognize the contribution of previous generations to inbreeding in the present study population. This requires accounting for “genetic sampling” (Weir, 1996) between generations, thereby leading to an ibd interpretation of inbreeding: ibd alleles descend from a single allele in a reference population. It also allows the prediction of inbreeding coefficients by path counting when pedigrees are known (Wright, 1922). If individual J is ancestral to both individuals (j^{prime}) and j″, and if there are n individuals in the pedigree path joining (j^{prime}) to j″ through J, then Fj = ∑(0.5)n(1 + FJ) where FJ is the inbreeding coefficient of ancestor J and Fj is the inbreeding coefficient of offspring j of parents (j^{prime}) and j″. The sum is over all ancestors J and all paths joining (j^{prime}) to j″ through J. The expression is also the coancestry ({theta }_{j^{prime} j^{primeprime} }) of (j^{prime}) and j″: the probability an allele drawn randomly from (j^{prime}) is ibd to an allele drawn randomly from j″.The allele proportion pl in a study population has expectation πl over evolutionary replicates of the population from an ancestral reference population to the present time. Sample allele proportions ({tilde{p}}_{l}) provide information about the population proportions pl, and their statistical sampling properties follow from the binomial distribution. We do not invoke a specific genetic sampling distribution for the pl about their expectations πl although we do assume the second moments of that distribution depend on probabilities of ibd for pairs of alleles. One consequence of the assumed moments is that the probability of individual j in the study sample being heterozygous, i.e., the total expected value ({{{{{{mathcal{E}}}}}}}_{T}) of the heterozygosity indicator over replicates of the history of that individual, is$${{{{{{mathcal{E}}}}}}}_{T}({tilde{H}}_{{j}_{l}})=2{pi }_{l}(1-{pi }_{l})(1-{F}_{j})$$
    (5)
    The quantity Fj is the individual-specific version of FIT of Wright (1922) and we can regard it as the probability the two alleles at any locus for individual j are ibd. There is an implicit assumption in Eq. (5) that the reference population needed to define ibd is infinite and in HWE: there is probability Fj that j has homologous alleles with a single ancestral allele in that population and probability (1 − Fj) of j having homologous alleles with distinct ancestral alleles there. In the first place, the single ancestral allele has probability π of being the reference allele for that locus and the implicit assumption is that two ancestral alleles are both the reference type with probability π2. This does not mean there is an actual ancestral population with those properties, any more than use of ({{{{{{mathcal{E}}}}}}}_{T}) means there are actual replicates of the history of any population or individual, and we note that Eq. (5) does not allow higher heterozygosity than predicted by HWE. Nonetheless, the concept of ibd allows theoretical constructions of great utility and we now present a framework for approaching empirical situations.Inbreeding, or ibd, implies a common ancestral origin for uniting alleles and statements about sample allele proportions ({tilde{p}}_{l}) require consideration of possible ibd for other pairs of alleles in the sample. The total expectation of (2{tilde{p}}_{l}(1-{tilde{p}}_{l})) over samples from the population and over evolutionary replicates of the study population is ((Weir, 1996), p176)$${{{{{{mathcal{E}}}}}}}_{T}[2{tilde{p}}_{l}(1-{tilde{p}}_{l})]=2{pi }_{l}(1-{pi }_{l})left[(1-{theta }_{S})-frac{1}{2n}left(1+{F}_{W}-2{theta }_{S}right)right]$$
    (6)
    where FW is the parametric inbreeding coefficient averaged over sample members, ({F}_{W}=mathop{sum }nolimits_{j = 1}^{n}{F}_{j}/n), and θS is the average parametric coancestry in the sample, ({theta }_{S}=mathop{sum }nolimits_{j = 1}^{n}{sum }_{j^{prime} ne j}{theta }_{jj^{prime} }/[n(n-1)]). Equivalent expressions were given by McPeek et al. (2004) and DeGiorgio and Rosenberg (2009). We note the relationship fW = (FW − θS)/(1 − θS) given by Wright (1922) and we showed in WG17 the equivalent expression fj = (Fj − θS)/(1 − θS) for individual-specific values (θS is Wright’s FST).For a large number of SNPs, the expectation of a ratio estimator of the type considered here is the ratio of expectations (Ochoa & Storey, 2021). Therefore, the total expectations of the ({hat{f}}_{{{{{{{rm{Hom}}}}}}}_{j}}), taking into account both statistical and genetic sampling, are$${{{{{{mathcal{E}}}}}}}_{T}({hat{f}}_{{{{{{{rm{HOM}}}}}}}_{j}})=1-frac{1-{F}_{j}}{(1-{theta }_{S})-frac{1}{2n}left(1+{F}_{W}-2{theta }_{S}right)}=frac{{f}_{j}-frac{1}{2n}(1+{f}_{W})}{1-frac{1}{2n}(1+{f}_{W})}$$
    (7)
    For all sample sizes, ({hat{f}}_{{{{{{{rm{HOM}}}}}}}_{j}}) has an expected value less than the true value fj, with the bias being of the order of 1/n. The ranking of ({{{{{{mathcal{E}}}}}}}_{T}({hat{f}}_{{{{{{{rm{HOM}}}}}}}_{j}})) values, however, is the same as the ranking of the fj and, therefore, of the Fj. For large sample sizes, Eq. (7) reduces to ({{{{{{mathcal{E}}}}}}}_{T}({hat{f}}_{{{{{{{rm{HOM}}}}}}}_{j}})={f}_{j}). Averaging over individuals shows that ({{{{{{mathcal{E}}}}}}}_{T}({hat{f}}_{{{{{{rm{HOM}}}}}}})={f}_{W}): the population-level estimator in Eq. (2) has total expectation of fW, not FW.A different outcome is found for the ({hat{f}}_{{{{{{{rm{UNI}}}}}}}_{j}}) estimator of Yengo et al. (2017) (i.e., ({hat{f}}^{III}) of Yang et al. (2011); ({hat{f}}_{{{{{{rm{GCTA}}}}}}3}) of (Gazal et al., 2014)). This estimator, with the weighted (w) ratio of averages over loci we recommend, as opposed to the unweighted (u) average of ratios over loci used in their papers, is$${hat{f}}_{{{{{{{rm{UNI}}}}}}}_{j}}^{w}=frac{mathop{sum }nolimits_{l = 1}^{L}[{X}_{jl}^{2}-(1+2{tilde{p}}_{l}){X}_{jl}+2{tilde{p}}_{l}^{2}]}{mathop{sum }nolimits_{l = 1}^{L}2{tilde{p}}_{l}(1-{tilde{p}}_{l})}$$
    (8)
    In this equation Xjl is the reference allele dosage, the number of copies of the reference allele, at SNP l for individual j. It is equivalent to the estimator given by (Ritland (1996), eq. 5) and attributed by him to Li & Horvitz (1953).Ochoa & Storey (2021) showed that ({hat{f}}_{{{{{{{rm{UNI}}}}}}}_{j}}^{w}) has expectation, for a large number of SNPs and a large sample size, of$${{{{{{mathcal{E}}}}}}}_{T}({hat{f}}_{{{{{{{rm{UNI}}}}}}}_{j}}^{w})=frac{{F}_{j}-2{{{Psi }}}_{j}+{theta }_{S}}{1-{theta }_{S}}={f}_{j}-2{psi }_{j}$$
    (9)
    where Ψj is the average coancestry of individual j with other members of the study sample: ({{{Psi }}}_{j}=mathop{sum }nolimits_{j^{prime} = 1,j^{prime} ne j}^{n}{theta }_{jj^{prime} }/(n-1)). We term ψj = (Ψj − θS)/(1 − θS) the within-population individual-specific average kinship coefficient. The Ψj have an average of θS over members of the sample, so the average of the ψj’s is zero and expected value of the average of the ({hat{f}}_{{{{{{{rm{UNI}}}}}}}_{j}}^{w}) is fW, as is the case for ({hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}) below.Equation (9) shows that the ({hat{f}}_{{{{{{{rm{UNI}}}}}}}_{j}}^{w}) have expected values with the same ranking as the Fj values only if every individual j in the sample has the same average kinship ψj with other sample members.Finally, we mention another common estimator described by VanRaden (2008), termed fGCTA1 by Gazal et al. (2014) and available from the GCTA software (Yang et al., 2011) with option –ibc. We referred to this as the “standard” estimator in WG17. The weighted version for multiple loci is$${hat{f}}_{{{{{{{rm{STD}}}}}}}_{j}}^{w}=frac{{sum }_{l}{({X}_{jl}-2{tilde{p}}_{l})}^{2}}{{sum }_{l}2{tilde{p}}_{l}(1-{tilde{p}}_{l})}-1$$
    (10)
    and it has the large-sample expectation of (fj − 4ψj) as is implied by WG17 (Eq. 13) and as was given by Ochoa & Storey (2021). We summarize the various measures of inbreeding and coancestry in Table 1, and we include sample sizes in the expectations shown in Table 2.Table 1 Measures of inbreeding and coancestry.Full size tableTable 2 Estimators of inbreeding.Full size tableThe ({hat{f}}_{{{{{{rm{HOM}}}}}}}), ({hat{f}}_{{{{{{rm{UNI}}}}}}},{hat{f}}_{{{{{{rm{STD}}}}}}}) and ({hat{f}}_{{{{{{rm{MLE}}}}}}}) estimators of individual or population inbreeding coefficients make explicit use of sample allele proportions. This means that all four have small-sample biases, and none of the four provide estimates of the ibd quantities F or Fj. We showed that ({hat{f}}_{{{{{{rm{HOM}}}}}}}) is actually estimating the within-population inbreeding coefficients: the total inbreeding coefficients relative to the average coancestry of pairs of individuals in the sample, but ({hat{f}}_{{{{{{rm{UNI}}}}}}}) and ({hat{f}}_{{{{{{rm{STD}}}}}}}) are estimating expressions that also involve average kinships ψ.Allele sharingIn a genetic sampling framework, and with the ibd viewpoint, we consider within-individual allele sharing proportions Ajl for SNP l in individual j (we wrote M rather than A in WG17 and in (Goudet et al., 2018)). These equal one for homozygotes and zero for heterozygotes and sample values can be expressed in terms of allele dosages, ({tilde{A}}_{jl}={({X}_{jl}-1)}^{2}). We also consider between-individual sharing proportions ({A}_{jj^{prime} l}) for SNP l and individuals j and (j^{prime}). These are equal to one for both individuals being the same homozygote, zero for different homozygotes, and 0.5 otherwise. Observed values can be written as ({tilde{A}}_{jj^{prime} l}=[1+({X}_{jl}-1)({X}_{j^{prime} l}-1)]/2), with an average over all pairs of distinct individuals in a sample of ({tilde{A}}_{Sl}). Astle & Balding (2009) introduced ({tilde{A}}_{jj^{prime} l}) as a measure of identity in state of alleles chosen randomly from individuals j and (j^{prime}), and Ochoa & Storey (2021) used a simple transformation of this quantity. The allele sharing for an individual with itself is Ajjl = (1 + Ajl)/2.The same logic that led to Eq. (5) provides total expectations for allele-sharing proportions for all (j,j^{prime}):$$begin{array}{lll}{{{{{{mathcal{E}}}}}}}_{T}({tilde{A}}_{jj^{prime} l})&=&1-2{pi }_{l}(1-{pi }_{l})(1-{theta }_{jj^{prime} })\ {{{{{{mathcal{E}}}}}}}_{T}({tilde{A}}_{Sl})&=&1-2{pi }_{l}(1-{pi }_{l})(1-{theta }_{S})end{array}$$Note that θjj = (1 + Fj)/2. The nuisance parameter 2πl(1 − πl) cancels out of the ratio ({{{{{{mathcal{E}}}}}}}_{T}({tilde{A}}_{jj^{prime} l}-{tilde{A}}_{Sl})/{{{{{{mathcal{E}}}}}}}_{T}(1-{tilde{A}}_{Sl})) and this motivates definitions of allele-sharing estimators of the inbreeding coefficient for individual j and the kinship coefficient for individuals (j,j^{prime}) as$${hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}=frac{{sum }_{l}({tilde{A}}_{{j}_{l}}-{tilde{A}}_{{S}_{l}})}{{sum }_{l}(1-{tilde{A}}_{Sl})},{hat{psi }}_{{{{{{{rm{AS}}}}}}}_{jj^{prime} }}=frac{{sum }_{l}({tilde{A}}_{jj^{prime} l}-{tilde{A}}_{{S}_{l}})}{{sum }_{l}(1-{tilde{A}}_{Sl})}$$
    (11)
    For a large number of SNPs, these are unbiased for fj and ({psi }_{jj^{prime} }) for all sample sizes. We showed in WG17 there is no need to filter on minor allele frequency to preserve the lack of bias. Note that ({hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}) is a linear function of the form ({a}_{S}+{b}_{S}{tilde{A}}_{j}) with ({tilde{A}}_{j}) being the total homozygosity for j and constants aS, bS being the same for all individuals j. Changing the scope of the study, from population to world for example, preserves linearity (with different values of aS, bS). The changed estimates are linear functions of the old estimates: old and new estimates are completely correlated and are rank invariant over all samples that include particular individuals, i.e., over all reference populations. Unlike the case for ({hat{f}}_{{{{{{rm{UNI}}}}}}}) or ({hat{f}}_{{{{{{rm{STD}}}}}}}), rank invariance is guaranteed for ({hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}) for any two individuals even if only one more individual is added to the study.For large sample sizes, ((1-{tilde{A}}_{Sl})approx 2{tilde{p}}_{l}(1-{tilde{p}}_{l})). Under that approximation, ({hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}) is the same as ({hat{f}}_{{{{{{{rm{Hom}}}}}}}_{j}}) but the approximation is not necessary in computer-based analyses. Summing the large-sample estimates over individuals not equal to j gives an estimator for the average individual kinship ψj:$${hat{psi }}_{{{{{{{rm{AS}}}}}}}_{j}}=-frac{{sum }_{l}({X}_{jl}-2{tilde{p}}_{l})(1-2{tilde{p}}_{l})}{{sum }_{l}4{tilde{p}}_{l}(1-{tilde{p}}_{l})}$$
    (12)
    Adding (2{hat{psi }}_{{{{{{{rm{AS}}}}}}}_{j}}) to ({hat{f}}_{{{{{{{rm{UNI}}}}}}}_{j}}^{w}) gives ({hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}), as expected, as does adding (4{hat{psi }}_{{{{{{{rm{AS}}}}}}}_{j}}) to ({hat{f}}_{{{{{{{rm{STD}}}}}}}_{j}}^{w}). Similarly, ({hat{psi }}_{{{{{{{rm{AS}}}}}}}_{jj^{prime} }}) is obtained by adding ({hat{psi }}_{{{{{{{rm{AS}}}}}}}_{j}}) and ({hat{psi }}_{{{{{{{rm{AS}}}}}}}_{j^{prime} }}) to ({hat{psi }}_{{{{{{{rm{STD}}}}}}}_{jj^{prime} }}), where (Yang et al., 2011)$${hat{psi }}_{{{{{{{rm{STD}}}}}}}_{jj^{prime} }}=frac{mathop{sum}nolimits_{l}({X}_{jl}-2{tilde{p}}_{l})({X}_{j^{prime} l}-2{tilde{p}}_{l})}{mathop{sum}nolimits_{l}4{tilde{p}}_{l}(1-{tilde{p}}_{l})}$$These are the elements of the first method for constructing the GRM given by VanRaden (2008).When inbreeding and coancestry coefficients are defined as ibd probabilities they are non-negative, but the within-population values f and ψ will be negative for individuals, or pairs of individuals, having smaller ibd allele probabilities than do pairs of individuals in the sample, on average. Individual-specific values of f always have the same ranking as the individual-specific F values, and they are estimable. Negative estimates can be avoided by the transformation to (({hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}-{hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}^{min })/(1-{hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}^{min })) where ({hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}}^{min }) is the smallest value over individuals of the ({hat{f}}_{{{{{{{rm{AS}}}}}}}_{j}})’s. We don’t see the need for this transformation, and we noted above the recognition of the utility of negative values. Ochoa & Storey (2021) wished to estimate Fj rather than fj and, to overcome the lack of information about the ancestral population serving as a reference point for ibd, they assumed the least related pair of individuals in a sample have a coancestry of zero. We showed in WG17 that this brings estimates in line with path-counting predicted values when founders are assumed to be not inbred and unrelated, but we prefer to avoid the assumption. We stress that, absent external information or assumptions, F is not estimable. Instead, linear functions of F that describe ibd of target pairs of alleles relative to ibd in a specified set of alleles are estimable and have utility in empirical studies.Runs of homozygosityEach of the inbreeding estimators considered so far has been constructed for individual SNPs and then combined over SNPs. Observed values of allelic state are used to make inferences about the unobserved state of identity by descent. Estimators based on ROH, however, suppose that ibd for a region of the genome can be observed. Although F is the probability an individual has ibd alleles at any single SNP, in fact ibd occurs in blocks within which there has been no recombination in the paths of descent from common ancestor to the individual’s parents. Whereas a single SNP can be homozygous without the two alleles being ibd, if many adjacent SNPs are homozygous the most likely explanation is that they are in a block of ibd (Gibson et al., 2006). There can be exceptions, from mutation for example, and several publications give strategies for identifying runs of homozygotes for which ibd may be assumed (e.g., Gazal et al. (2014); (Joshi et al., 2015)). These strategies include adjusting the size of the blocks, the numbers of heterozygotes or missing values allowed per block, the minor allele frequency, and so on. These software parameters affect the size of the estimates (Meyermans et al., 2020). Some methods (e.g., Gazal et al. (2014); (Narasimhan et al., 2016)) use hidden Markov models where ibd is the hidden status of an observed homozygote. Model-based approaches necessarily have assumptions, such as HWE in the sampled population.We provide more details elsewhere, but we note here that ROH methods offer a useful alternative to SNP-by-SNP methods even though they cannot completely compensate for lack of information on the ibd reference population. We note also that shorter runs of ibd result from more distant relatedness of an individual’s parents, and ROH procedures can be set to distinguish recent (familial) ibd from distant (evolutionary) ibd. SNP-by-SNP estimators do not make a distinction between these two time scales. More

  • in

    Species richness and identity both determine the biomass of global reef fish communities

    Reef life surveyReef fish communities were censused by a combination of experienced marine scientists and trained recreational SCUBA divers using globally standardized Reef Life Survey methods. All surveys were undertaken on 50 m long transects laid along a contour (at consistent depth) on predominantly hard substrate (usually rocky or coral reef) in shallow waters (depth range of transects 1 to 20 m, average ~7.2 m). Full details of fish census methods, data quality, and training of divers are provided in refs. 22,34,35 and in an online methods manual (www.reeflifesurvey.com). Fish abundance counts and size estimates per 500 m2 transect area (2 ×250 m2 blocks) were converted to biomass using length–weight relationships for each species obtained from Fishbase (www.fishbase.org). In cases where length–weight relationships were provided in Fishbase using standard length or fork length, rather than total length as estimated by divers, length–length relationships provided in Fishbase allowed conversion to the total length. For improved accuracy in biomass assessments, observed sizes were also adjusted to account for the bias in divers’ perception of fish size underwater using an empirical calibration36. Length–weight coefficients from similar-shaped close relatives were used for those species where length–weight relationships were not available in Fishbase. All transects were collapsed into a single average value of biomass for each species at a location to account for any differences in the total number of transect surveys performed.Decomposition of difference in ecosystem functioningOur equation was inspired by previous decompositions, principally the Price equation originally derived in the field of evolutionary biology as a means of separating genetic and environmental influences on phenotypic change over time37. Fox38 and later Fox and Kerr12 modified the Price equation to describe how the difference in the ecological function between two communities can be decomposed into components with different ecological interpretations. We follow a similar approach but use a different decomposition where the resulting components are similar to, but not the same as, the components proposed by Fox and Kerr12.We begin by assuming that the ecological function of the community, such as biomass, is a simple additive function of the contributions of its constituent species. We go on to compare two communities, one of which we consider the “reference” community and the other we refer to as the “comparison” community. The species present in the reference community can be classified into two types: species that are unique to the reference community (i.e., not present in the comparison community) and those that are in common with the comparison community. Let suB be the number of unique species in the reference community, and sc be the number in common between the two communities. Let ({bar{z}}_{{uB}}) be the average ecological function contributed per unique species to the reference community, and ({bar{z}}_{{cB}}) be the average ecological function contributed per shared species in the reference community. The total ecological function TB of the reference community can thus be decomposed as:$${T}_{B}={s}_{{uB}}{bar{z}}_{{uB}}+{s}_{c}{bar{z}}_{{cB}}$$
    (1)
    where the first term represents the ecological function contributed by species that are unique to the reference community (i.e., not present in the comparison community) and the latter term represents the contribution from species that are also found in the comparison community.Analogously, in the comparison community, the total ecological function can be decomposed as:$${T}_{F}={s}_{{uF}}{bar{z}}_{{uF}}+{s}_{c}{bar{z}}_{{cF}}$$
    (2)
    with a similar interpretation to Eq. (1). Though there are sc species in common between the two communities, the average per species contribution need not be the same in the two communities (i.e., ({bar{z}}_{{cB}}) may differ from ({bar{z}}_{{cF}})).The species in common between the two communities can serve as a reference point for comparison between communities. It is useful to define ({delta }_{B}={bar{z}}_{{uB}}-{bar{z}}_{{cB}}) and ({delta }_{F}={bar{z}}_{{uF}}-{bar{z}}_{{cF}}) as the difference in average ecological function per species of unique species versus shared species in reference and comparison communities, respectively. From this perspective, we consider the average ecological function of a species unique to the reference community as being equal to the average ecological function of shared species (as measured in the same community) plus the deviation from this value ({bar{z}}_{{uB}}={bar{z}}_{{cB}}+{delta }_{B}). Using this equality and the analogous one for ({bar{z}}_{{uF}}), along with Eqs. (1) and (2), the difference in the ecological function between communities can be decomposed as$$Delta T={T}_{F}-{T}_{B}={-s}_{{uB}}{bar{z}}_{{cB}}-{s}_{{uB}}{delta }_{B}+{s}_{{uF}}{bar{z}}_{{cF}}+{s}_{{uF}}{delta }_{F}+{s}_{c}left({bar{z}}_{{cF}}-{bar{z}}_{{cB}}right)$$
    (3)
    The first two terms represent the loss in ecological function in the comparison community due to the loss of species that are unique to the reference community. Specifically, the first term represents the loss in ecological function due to the absence of unique species if these species had the same average value of functioning as each of the shared species. In other words, it is the amount by which biomass is expected to decline if species were interchangeable. Therefore, we interpret this term as the “richness loss” or the loss in functioning due strictly to the loss of species: RICH-L ((={-s}_{{uB}}{bar{z}}_{{cB}})). It will always be negative, assuming there is at least one species unique to the reference population. In cases where ({bar{z}}_{{cB}} > {bar{z}}_{{uB}}), it is possible for RICH-L to exceed the total functioning observed at the reference site, which complicates interpretation of the raw values. In this case, it is useful to consider only the relative quantities (each component is scaled by the sum of the absolute values of all components). We note that this situation arises only 41 times out of 2867 comparisons in our analysis, and removing these cases has no effect on our findings. We advise future applications be aware of this potential issue and test for its influence.The second term accounts for the fact that the true loss in ecological function due to these lost species will often differ from the “richness expectation” because the lost species differ in value from the average value of shared species. In other words, this term reflects the deviation in the actual contributions of lost species from the average of shared species, which implies that not all species contribute equally (and that the identities of the species are important in determining differences in biomass between the two communities). We, therefore, interpret this term as indicating “compositional loss,” or the degree to which loss in biomass is due to loss of particular species: COMP-L ((= – {s}_{{uB}}{delta}_{B})). If the average lost species provide a higher contribution to the reference community than the average shared species (({bar{z}}_{{uB}} > {bar{z}}_{{cB}})), the COMP-L term will be negative. On the other hand, if the average lost species represent lower contributions, the COMP-L term will be positive (({bar{z}}_{{uB}} < {bar{z}}_{{cB}})).The next two terms are analogous to the first two terms but instead represent the increase in ecological function in the comparison community due to the “gain” of unique species that are lacking from the reference community. The third term represents the expected increase in ecological function due to an increase in species richness assuming these gained species had the same per species contribution as the shared species: RICH-G ((={+s}_{{uF}}{bar{z}}_{{cF}})). It is always positive, assuming the comparison community has at least one unique species. The fourth term, COMP-G ((=+{s}_{{uF}}{delta }_{F})), reflects the difference in composition (with respect to average value) of gained versus shared species. This term can be positive or negative, being positive if the gained species have a higher per species value than the shared species.The final term focuses on the changes in biomass considering only the species that are present in both communities. This can be thought of as holding richness and composition constant and considering changes in the community biomass that are controlled extrinsically, i.e., by underlying gradients in resource availability and other environmental factors. Historically, this term has been referred to as the “context-dependent effect,” or CDE, and is the number of shared species (({s}_{c})), multiplied by the difference in biomasses among shared species at both sites ((={s}_{c}({bar{z}}_{{cF}}-{bar{z}}_{{cB}}))). It can be of either sign: positive if shared species have a higher value in the comparison community than in the reference, negative if they have a higher value in the reference community. The number of shared species has the potential to bias away from the CDE term if it is very low. However, we note that, on average, 49.1 ± 0.003% of species are shared for each comparison at the 100-km scale, and this value is remarkably consistent regardless of spatial scale (51.3–50.0% for 15–50 km).Our decomposition is similar to, but not the same as, that of Fox and Kerr12, though both are mathematically sound. Only the CDE term is mathematically identical across the two decompositions and, thus, shares the same interpretation. By extension, the sum across the loss and gain terms (the total diversity effect, or DIV) must also be identical, because both equations partition the same total quantity. Thus, it is important to note that using either decomposition yields the same inference with respect to comparisons of DIV and CDE.Our decomposition differs from Fox and Kerr’s because the two approaches use different reference points. We take the perspective that the shared species form the basis for comparison between two communities, so we then evaluate the average value of a unique species with respect to its deviation from an average value of a shared species. In contrast, Fox and Kerr effectively evaluate the average value of a unique species with respect to its deviation from the average value of any species in that community (averaging over both unique and shared species). In both decompositions, the “composition” components only exist if there is some difference in the average value of shared and unique species. We prefer our decomposition for this case because it works with that difference directly rather than indirectly via the difference between unique and all species (which is the average of unique and shared species). Moreover, our composition makes intuitive sense that the function of the “average” species is determined by the ones that are known to exist at both sites. A full comparison of the Fox and Kerr formulation and ours is provided in the Supplementary Materials.Statistical analysisA general function to conduct our new decomposition from a site-by-species biomass matrix, and a second function to perform the simulations, can be found here: https://gist.github.com/jslefche/76c076c1c7c5d200e5cb87113cdb9fb4.We first ordered all sites by decreasing total biomass. Beginning with the highest biomass site of all sites as the first reference site, we identified all other sites within a certain spatial radius (15-, 25-, 50-, or 100-km) to serve as the comparison sites. Setting the reference to be the site with the highest community biomass constrains the sum of the terms to be negative. This choice simplifies the language used to discuss the output13 and allows us to speak directly to the consequences of real-world activities like overharvesting (and their implications).We then computed the components for each set of comparisons. We standardized the output to the same scale (−1, 1) by first taking the sum of the absolute value of all components, and then dividing each component by this value. This relativization was done to account for the fact that raw biomass may differ substantially among sites and regions and to make our results comparable across the entire dataset. Once the scaled components were computed, the reference and comparison sites were removed from the ordered list from any further comparisons to prevent any bias that might arise from including the same site multiple times. We then moved onto the next most productive site in the list, identified the comparison sites within 100 km, computed the components, and so on, until all sites were analyzed. From these individual comparisons, we computed the means of all components while omitting any reference sites for which there were fewer than five comparison sites. We alternately averaged the components for all comparisons for each reference site and then took the grand mean of these averaged values, although this additional level of aggregation did not qualitatively change our results (Supplementary Fig. 6). We have chosen to present the raw values in the main text to demonstrate the full range of variability inherent in the individual comparisons, which might otherwise be condensed by showing only the means for each reference site. We repeated the analysis over multiple spatial radii to assess whether the spatial extent and therefore the size and composition of the species pool, might influence our results.We calculated the relative strength of the total diversity effect vs. the context-dependent effect for each comparison as the ratio of DIV/CDE, and of compositional vs. richness losses as:$${{{{{rm{Q}}}}}}=frac{(-{s}_{{uB}}{delta }_{B}{-s}_{{uB}}{bar{z}}_{{cB}})}{{-s}_{{uB}}{bar{z}}_{{cB}}}=frac{{bar{z}}_{{uB}}}{{bar{z}}_{{cB}}}$$ (4) In this case, Q = (COMP-L + RICH-L)/RICH-L, which reduces to the average value of unique species relative to the average value of shared species at the reference site. This quantity reflects the magnitude to which species unique to the reference site contribute to biomass relative to the “expected” contribution per species. To avoid biases associated with averaging ratios, we report the geometric mean of both quantities. Bootstrapped 95% confidence intervals were derived by randomly resampling DIV/CDE and Q for a total of 5000 times. For DIV/CDE, some values were negative, so we excluded them in both the original data and bootstrap samples. As an alternative approach that focused on the magnitude of effect, we examined the absolute value of |DIV | / | CDE | . In this case, the ratio was 6.9x with bootstrap 95% CIs of [6.2, 7.7].To explore the drivers of the components of our decomposition, we applied random forest analysis to account for potential collinearity and interactions among the suite of predictors previously selected in ref. 39. Depth was recorded on the surveys while the following predictors were obtained from the combination of remote sensed and in situ measurements compiled in the Bio-ORACLE database: mean, minimum, maximum, and range of sea surface temperature; mean, minimum and maximum for surface chlorophyll-a; mean salinity; mean PAR; mean dissolved oxygen; mean nitrate concentration; mean phosphate concentration40. Finally, an index of human population density was calculated by fitting a smoothly tapered surface to each settlement point on the year 2010 world-population density grid using a quadratic kernel function described previously41. Random forests were fit using the default settings in the randomForest package42 in R version 4.1.143. Variable importance was determined using the percent increase in the mean-square error after randomly permuting the predictor of interest for each tree in the random forest, averaging the error of the models, and then computing the difference relative to the accuracy of the original model.Null simulationsA key finding of our analysis is that compositional losses are considerably greater than losses due to other aspects of the reef fish community. We wanted to evaluate the possibility of whether such a result could be an artifact of applying our decomposition to a dataset in which we assign the site with the higher total biomass as the “reference” community and the site with lower total biomass as the “focal” community. To do so, we conducted simulations in which we created communities with species richness values matching the observed data, but for community compositions that were random. Following the same procedure we used with the real communities, we applied our decomposition to these simulated communities to generate null distributions for the average values of each of the five terms when community composition is random. Comparing our observed values to these null distributions tells us if the values of the compositional components (or indeed any component) we observed arose as an artifact of our procedure or, alternatively, because high-biomass sites actually contain more high-biomass species than expected under random community assembly.Our simulation procedure focused on the site-by-species biomass matrix from each set of comparisons used in the main 100-km analysis. We divided this matrix by the corresponding site-by-species abundance matrix to yield the observed per capita contribution of each species in each community. We then averaged the per capita contributions of each species across all communities where the species was present to yield a single vector representing mean per capita contributions for all S species within that set of comparisons.We initially constructed each simulated community by populating it with every species in the region (“maximum richness”). To determine the biomass of each species in each community we applied the following procedure. First, we identified the minimum and maximum observed abundance of each species across all communities where it is present. For a single community, we sampled an integer value between the minimum and maximum abundance for each species to yield a single vector of random abundance values of length S, and then multiplied this vector by the vector of average per capita contributions. This procedure yielded a new vector representing a new total contribution to biomass by every species. We repeated this for all n communities in the original site-by-species matrix and bound these vectors together in a new “maximum richness” version of the site-by-species matrix. For the ith row (community) in the original dataset, we calculated the richness, si. We then randomly subsampled si species at random from the simulated “maximum richness” site-by-species matrix and set the biomass of any remaining species to zero. We repeated this for each community to yield a simulated “observed richness” site-by-species matrix with the same dimensions as the original matrix. This procedure ensures that richness is held at the observed levels and that the biomass contribution of each species are within the observed range.These communities were intentionally constructed randomly with respect to composition as our goal was to test whether the observed compositional effects in the real data are significantly different than under this null hypothesis with respect to composition. Thus, using the simulated “observed richness” site-by-species matrix, we computed the (scaled) components as we had with the real data and took their means across all communities. We repeated the randomization procedure 1000 times to yield 1000 total average values of each component. We compared the observed mean to the distribution of expected means using a one-tailed t-test to determine whether the observed components were more or less extreme than would be expected by chance.Reporting SummaryFurther information on research design is available in the Nature Research Reporting Summary linked to this article. More