More stories

  • in

    Opportunities and challenges of macrogenetic studies

    1.Brown, J. H. & Maurer, B. A. Macroecology: the division of food and space among species on continents. Science 243, 1145–1150 (1989).CAS 
    Article 

    Google Scholar 
    2.Gaston, K. J., Robinson, D. & Chown, S. L. Macrophysiology: large-scale patterns in physiological traits and their ecological implications. Funct. Ecol. 18, 159–167 (2004).Article 

    Google Scholar 
    3.Chown, S. L. & Gaston, K. J. Macrophysiology–progress and prospects. Funct. Ecol. 30, 330–344 (2016).Article 

    Google Scholar 
    4.Avise, J. C. Phylogeography: the History and Formation of Species (Harvard University Press, 2000).5.Ebach, M. C. Origins of Biogeography. Vol. 13 (Springer, 2015).6.Brundin, L. On the real nature of transantarctic relationships. Evolution 19, 496–505 (1965).
    Google Scholar 
    7.Beheregaray, L. B. Twenty years of phylogeography: the state of the field and the challenges for the Southern Hemisphere. Mol. Ecol. 17, 3754–3774 (2008).PubMed 
    PubMed Central 

    Google Scholar 
    8.Hickerson, M. J. et al. Phylogeography’s past, present, and future: 10 years after Avise, 2000. Mol. Phylogenet. Evol. 54, 291–301 (2010).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    9.Gaston, K. J. & Blackburn, T. M. A critique for macroecology. Oikos 84, 353–368 (1999).Article 

    Google Scholar 
    10.Lovegrove, B. G. The zoogeography of mammalian basal metabolic rate. Am. Nat. 156, 201–219 (2000).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    11.Reich, P. B., Walters, M. B. & Ellsworth, D. S. From tropics to tundra: Global convergence in plant functioning. Proc. Natl Acad. Sci. USA 94, 13730–13734 (1997).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    12.Chown, S. L. & Gaston, K. J. Macrophysiology for a changing world. Proc. Biol. Sci. 275, 1469–1478 (2008).PubMed 
    PubMed Central 

    Google Scholar 
    13.Kerr, J. T., Kharouba, H. M. & Currie, D. J. The macroecological contribution to global change solutions. Science 316, 1581–1584 (2007).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    14.Blanchet, S., Prunier, J. G. & De Kort, H. Time to go bigger: Emerging patterns in macrogenetics. Trends Genet. 33, 579–580 (2017). This study coined the term ‘macrogenetics’ and illustrated, through three study examples, how shifting toward macrogenetics should generate new perspectives and theories concerning genetic diversity patterns.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    15.Blanchet, S. et al. A river runs through it: the causes, consequences, and management of intraspecific diversity in river networks. Evol. Appl. 13, 1195–1213 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    16.Frankham, R. Resolving conceptual issues in conservation genetics: the roles of laboratory species and meta-analyses. Hereditas 130, 195–201 (2004).Article 

    Google Scholar 
    17.Arnqvist, G. & Wooster, D. Meta-analysis: synthesizing research findings in ecology and evolution. Trends Ecol. Evol. 10, 236–240 (1995).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    18.Paz-Vinas, I. et al. Systematic conservation planning for intraspecific genetic diversity. Proc. Biol. Sci. 285, 20172746 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    19.Pelletier, T. A. & Carstens, B. C. Geographical range size and latitude predict population genetic structure in a global survey. Biol. Lett. 14, 20170566 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    20.Miraldo, A. et al. An anthropocene map of genetic diversity. Science 353, 1532–1535 (2016). This paper is thought to be the first published study to massively repurpose public mtDNA sequences to explore global genetic patterns (100,791 sequences from >4,500 terrestrial mammal and amphibian species).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Yiming, L. et al. Latitudinal gradients in genetic diversity and natural selection at a highly adaptive gene in terrestrial mammals. Ecography 44, 206–218 (2021). This study found that adaptive IGV is higher at low latitudes and in smaller mammal species using repurposed MHC gene data from 93 mammal species.Article 

    Google Scholar 
    22.Manel, S. et al. Global determinants of freshwater and marine fish genetic diversity. Nat. Commun. 11, 692 (2020). This study repurposed 58,565 public mtDNA sequences from 5,912 freshwater and marine fish to explore the effects of environmental drivers (temperature, species diversity) on intraspecific genetic diversity.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    23.Theodoridis, S. et al. Evolutionary history and past climate change shape the distribution of genetic diversity in terrestrial mammals. Nat. Commun. 11, 2557 (2020). This study revealed a negative effect of past rapid climate change and a positive effect of interannual precipitation variability in shaping the genetic diversity of terrestrial mammals using 46,965 mtDNA sequences.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    24.Barrow, L. N., da Fonseca, E. M., Thompson, C. E. P. & Carstens, B. C. Predicting amphibian intraspecific diversity with machine learning: Challenges and prospects for integrating traits, geography, and genetic data. Mol. Ecol. Resour. https://doi.org/10.1111/1755-0998.13303 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    25.De Kort, H. et al. Life history, climate and biogeography interactively affect worldwide genetic diversity of plant and animal populations. Nat. Commun. 12, 516 (2021). This study found weak support for latitudinal IGV gradients, taxonomic-specific effects of temperature stability and life-history traits, and higher IGV in animals compared to plants using microsatellite and amplified fragment length polymorphism data from 8,386 local populations from 727 animal and plant species.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    26.Schmidt, C., Domaratzki, M., Kinnunen, R. P., Bowman, J. & Garroway, C. J. Continent-wide effects of urbanization on bird and mammal genetic diversity. Proc. Biol. Sci. 287, 20192497 (2020). This study used archived microsatellite data from 85 studies (66 species) to explore the effects of urbanization in mammals and birds.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    27.Millette, K. L. et al. No consistent effects of humans on animal genetic diversity worldwide. Ecol. Lett. 23, 55–67 (2020). The authors of this article conducted spatial and temporal analysis of the effects of humans on animal genetic diversity worldwide, by repurposing 175,247 mtDNA sequences from >17,000 animal species.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    28.Taberlet, P. et al. Genetic diversity in widespread species is not congruent with species richness in alpine plant communities. Ecol. Lett. 15, 1439–1448 (2012). This paper reports a Class I macrogenetic study based on amplified fragment length polymorphism genetic data from 27 alpine plant species that tested whether genetic and species diversities co-vary.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    29.Manel, S. et al. Broad-scale adaptive genetic variation in alpine plants is driven by temperature and precipitation. Mol. Ecol. 21, 3729–3738 (2012).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    30.Gugerli, F. et al. Relationships among levels of biodiversity and the relevance of intraspecific diversity in conservation – a project synopsis. Perspect. Plant. Ecol. Evol. Syst. 10, 259–281 (2008).Article 

    Google Scholar 
    31.Schlaepfer, D. R., Braschler, B., Rusterholz, H.-P. & Baur, B. Genetic effects of anthropogenic habitat fragmentation on remnant animal and plant populations: a meta-analysis. Ecosphere 9, e02488 (2018).Article 

    Google Scholar 
    32.González, A. V., Gómez-Silva, V., Ramírez, M. J. & Fontúrbel, F. E. Meta-analysis of the differential effects of habitat fragmentation and degradation on plant genetic diversity. Conserv. Biol. 34, 711–720 (2020).PubMed 
    Article 

    Google Scholar 
    33.Ratnasingham, S. & Hebert, P. D. N. Bold: the barcode of life data system. Mol. Ecol. Notes 7, 355–364 (2007).CAS 
    PubMed 
    PubMed Central 
    Article 

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

    Google Scholar 
    35.Kattge, J. et al. TRY plant trait database–enhanced coverage and open access. Glob. Change Biol. 26, 119–188 (2020).Article 

    Google Scholar 
    36.Theodoridis, S., Rahbek, C. & Nogues-Bravo, D. Exposure of mammal genetic diversity to mid-21st century global change. Ecography 44, 817–831 (2021).Article 

    Google Scholar 
    37.Rissler, L. J. Union of phylogeography and landscape genetics. Proc. Natl Acad. Sci. USA 113, 8079–8086 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    38.Hubbell, S. P. The unified neutral theory of biodiversity and biogeography (Princeton University Press, 2001).39.Haldane, J. B. S. A mathematical theory of natural and artificial selection, Part V: selection and mutation. Math. Proc. Camb. Philos. Soc. 23, 838–844 (1927).Article 

    Google Scholar 
    40.Wright, S. Evolution in Mendelian populations. Genetics 16, 97–159 (1931).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    41.Fisher, R. A. On the dominance ratio. Proc. R. Soc. Edinburgh 42, 321–341 (1922).Article 

    Google Scholar 
    42.Kimura, M. & Weiss, G. H. The stepping stone model of population structure and the decrease of genetic correlation with distance. Genetics 49, 561–576 (1964).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    43.Kingman, J. F. C. The coalescent. Stoch. Process. Their Appl. 13, 235–248 (1982).Article 

    Google Scholar 
    44.Kimura, M. Evolutionary rate at the molecular level. Nature 217, 624–626 (1968).CAS 
    PubMed 
    Article 

    Google Scholar 
    45.Soulé, M. E. in Molecular Evolution (ed. Ayala, F. J.) 60–77 (Sinauer Associates, 1976).46.Brown, A. H. Isozymes, plant population genetic structure and genetic conservation. Tag. Theor. Appl. Genet. Theor. Angew. Genet. 52, 145–157 (1978).CAS 
    Article 

    Google Scholar 
    47.Mullis, K. et al. Specific enzymatic amplification of DNA in vitro: the polymerase chain reaction. Cold Spring Harb. Symp. Quant. Biol. 51, 263–273 (1986).CAS 
    PubMed 
    Article 

    Google Scholar 
    48.Sanger, F., Nicklen, S. & Coulson, A. R. DNA sequencing with chain-terminating inhibitors. Proc. Natl Acad. Sci. USA 74, 5463–5467 (1977).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    49.Miller, M. R., Dunham, J. P., Amores, A., Cresko, W. A. & Johnson, E. A. Rapid and cost-effective polymorphism identification and genotyping using restriction site associated DNA (RAD) markers. Genome Res. 17, 240–248 (2007).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    50.Carroll, E. L. et al. Genetic and genomic monitoring with minimally invasive sampling methods. Evol. Appl. 11, 1094–1119 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    51.Hebert, P. D. N., Cywinska, A., Ball, S. L. & deWaard, J. R. Biological identifications through DNA barcodes. Proc. Biol. Sci. 270, 313–321 (2003).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    52.Taberlet, P., Coissac, E., Pompanon, F., Brochmann, C. & Willerslev, E. Towards next-generation biodiversity assessment using DNA metabarcoding. Mol. Ecol. 21, 2045–2050 (2012).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    53.Gauthier, J. et al. Museomics identifies genetic erosion in two butterfly species across the 20th century in Finland. Mol. Ecol. Resour. 20, 1191–1205 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    54.Wandeler, P., Hoeck, P. E. A. & Keller, L. F. Back to the future: museum specimens in population genetics. Trends Ecol. Evol. 22, 634–642 (2007).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    55.Strasser, B. J. The experimenter’s museum: GenBank, natural history, and the moral economies of biomedicine. Isis 102, 60–96 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    56.Whitlock, M. C. Data archiving in ecology and evolution: best practices. Trends Ecol. Evol. 26, 61–65 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    57.Wilkinson, M. D. et al. The FAIR guiding principles for scientific data management and stewardship. Sci. Data 3, 160018 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    58.Deck, J. et al. The Genomic Observatories Metadatabase (GeOMe): A new repository for field and sampling event metadata associated with genetic samples. PLoS Biol. 15, e2002925 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    59.R Core Team. R: a language and environment for statistical computing, R Foundation for Statistical Computing http://www.r-project.org/index.html (2021).60.Manel, S. & Holderegger, R. Ten years of landscape genetics. Trends Ecol. Evol. 28, 614–621 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    61.Prunier, J. G., Colyn, M., Legendre, X., Nimon, K. F. & Flamand, M. C. Multicollinearity in spatial genetics: separating the wheat from the chaff using commonality analyses. Mol. Ecol. 24, 263–283 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    62.Stanley, R. R. E. et al. A climate-associated multispecies cryptic cline in the northwest Atlantic. Sci. Adv. 4, eaaq0929 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    63.Fenderson, L. E., Kovach, A. I. & Llamas, B. Spatiotemporal landscape genetics: investigating ecology and evolution through space and time. Mol. Ecol. 29, 218–246 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    64.Daza, J. M., Castoe, T. A. & Parkinson, C. L. Using regional comparative phylogeographic data from snake lineages to infer historical processes in middle America. Ecography 33, 343–354 (2010).
    Google Scholar 
    65.Riddle, B. R. Comparative phylogeography clarifies the complexity and problems of continental distribution that drove A. R. Wallace to favor islands. Proc. Natl Acad. Sci. USA 113, 7970–7977 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    66.Carstens, B. C., Morales, A. E., Field, K. & Pelletier, T. A. A global analysis of bats using automated comparative phylogeography uncovers a surprising impact of Pleistocene glaciation. J. Biogeogr. 45, 1795–1805 (2018).Article 

    Google Scholar 
    67.Smith, B. T., Seeholzer, G. F., Harvey, M. G., Cuervo, A. M. & Brumfield, R. T. A latitudinal phylogeographic diversity gradient in birds. PLoS Biol. 15, e2001073 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    68.Smith, B. T. et al. The drivers of tropical speciation. Nature 515, 406–409 (2014).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    69.Ballin, M., Barcaroli, G., Masselli, M. & Scarnó, M. Redesign Sample for Land Use/Cover Area Frame Survey (LUCAS) 2018 (EU Publications, 2018).70.Buchhorn, M. et al. Copernicus global land cover layers — Collection 2. Remote. Sens. 12, 1044 (2020).Article 

    Google Scholar 
    71.Jones, K. E. et al. PanTHERIA: a species-level database of life history, ecology, and geography of extant and recently extinct mammals: Ecological Archives E090-184. Ecology 90, 2648–2648 (2009).Article 

    Google Scholar 
    72.Tedesco, P. A. et al. A global database on freshwater fish species occurrence in drainage basins. Sci. Data 4, 170141 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    73.Vellend, M. & Geber, M. A. Connections between species diversity and genetic diversity: species diversity and genetic diversity. Ecol. Lett. 8, 767–781 (2005).Article 

    Google Scholar 
    74.Fourtune, L., Paz-Vinas, I., Loot, G., Prunier, J. G. & Blanchet, S. Lessons from the fish: a multi-species analysis reveals common processes underlying similar species-genetic diversity correlations. Freshw. Biol. 61, 1830–1845 (2016).Article 

    Google Scholar 
    75.Bertin, A. et al. Genetic variation of loci potentially under selection confounds species-genetic diversity correlations in a fragmented habitat. Mol. Ecol. 26, 431–443 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    76.Lawrence, E. R. & Fraser, D. J. Latitudinal biodiversity gradients at three levels: linking species richness, population richness and genetic diversity. Glob. Ecol. Biogeogr. 29, 770–788 (2020).Article 

    Google Scholar 
    77.Schmidt, C., Dray, S. & Garroway, C. J. Genetic and species-level biodiversity patterns are linked by demography and ecological opportunity. bioRxiv https://doi.org/10.1101/2020.06.03.132092 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    78.Hillebrand, H. On the generality of the latitudinal diversity gradient. Am. Nat. 163, 192–211 (2004).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    79.Pontarp, M. et al. The latitudinal diversity gradient: novel understanding through mechanistic eco-evolutionary models. Trends Ecol. Evol. 34, 211–223 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    80.Toews, D. P. L. & Brelsford, A. The biogeography of mitochondrial and nuclear discordance in animals. Mol. Ecol. 21, 3907–3930 (2012).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    81.Schmidt, C. & Garroway, C. J. The conservation utility of mitochondrial genetic diversity in macrogenetic research. Conserv. Genet. 22, 323–327 (2021).Article 

    Google Scholar 
    82.Gratton, P. et al. Which latitudinal gradients for genetic diversity? Trends Ecol. Evol. 32, 724–726 (2017). This response to Miraldo et al.20 identified a limitation of that article in that it did not account for the decay of genetic similarity with distance and represents the first critique of the downsides of the macrogenetic approach and the need for rigorous statistics.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    83.Loveless, M. D. & Hamrick, J. L. Ecological determinants of genetic structure in plant populations. Annu. Rev. Ecol. Syst. 15, 65–95 (1984).Article 

    Google Scholar 
    84.Hu, Y. et al. Spatial patterns and conservation of genetic and phylogenetic diversity of wildlife in China. Sci. Adv. 7, eabd5725 (2021).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    85.Johnson, M. T. J. & Munshi-South, J. Evolution of life in urban environments. Science 358, eaam8327 (2017).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    86.Aguilar, R., Quesada, M., Ashworth, L., Herrerias-Diego, Y. & Lobo, J. Genetic consequences of habitat fragmentation in plant populations: susceptible signals in plant traits and methodological approaches. Mol. Ecol. 17, 5177–5188 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    87.Pinsky, M. L. & Palumbi, S. R. Meta-analysis reveals lower genetic diversity in overfished populations. Mol. Ecol. 23, 29–39 (2014).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    88.Leigh, D. M., Hendry, A. P., Vázquez-Domínguez, E. & Friesen, V. L. Estimated six per cent loss of genetic variation in wild populations since the industrial revolution. Evol. Appl. 12, 1505–1512 (2019). This study estimated the magnitude of the loss of genetic variation over a century-scale using microsatellite data from 91 species.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    89.Schmidt, C. & Garroway, C. J. The population genetics of urban and rural amphibians in north America. Mol. Ecol. https://doi.org/10.1111/mec.16005 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    90.Bazin, E., Glémin, S. & Galtier, N. Population size does not influence mitochondrial genetic diversity in animals. Science 312, 570–572 (2006).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    91.Galtier, N., Nabholz, B., Glémin, S. & Hurst, G. D. D. Mitochondrial DNA as a marker of molecular diversity: a reappraisal. Mol. Ecol. 18, 4541–4550 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    92.Allio, R., Donega, S., Galtier, N. & Nabholz, B. Large variation in the ratio of mitochondrial to nuclear mutation rate across animals: implications for genetic diversity and the use of mitochondrial DNA as a molecular marker. Mol. Biol. Evol. 34, 2762–2772 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    93.Almeida-Rocha, J. M., Soares, L. A. S. S., Andrade, E. R., Gaiotto, F. A. & Cazetta, E. The impact of anthropogenic disturbances on the genetic diversity of terrestrial species: a global meta-analysis. Mol. Ecol. 29, 4812–4822 (2020).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    94.Landguth, E. L. et al. Quantifying the lag time to detect barriers in landscape genetics. Mol. Ecol. 19, 4179–4191 (2010).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    95.Paz-Vinas, I. et al. Macrogenetic studies must not ignore limitations of genetic markers and scale. Ecol. Lett. 24, 1282–1284 (2021).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    96.Crandall, E. D. et al. The molecular biogeography of the Indo-Pacific: testing hypotheses with multispecies genetic patterns. Glob. Ecol. Biogeogr. 28, 943–960 (2019).Article 

    Google Scholar 
    97.Excoffier, L. & Foll, M. fastsimcoal: a continuous-time coalescent simulator of genomic diversity under arbitrarily complex evolutionary scenarios. Bioinformatics 27, 1332–1334 (2011).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    98.Guillaume, F. & Rougemont, J. Nemo: an evolutionary and population genetics programming framework. Bioinformatics 22, 2556–2557 (2006).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    99.Phillips, J. D., French, S. H., Hanner, R. H. & Gillis, D. J. HACSim: an R package to estimate intraspecific sample sizes for genetic diversity assessment using haplotype accumulation curves. PeerJ Comput. Sci. 6, e243 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    100.Gratton, P. et al. A world of sequences: can we use georeferenced nucleotide databases for a robust automated phylogeography? J. Biogeogr. 44, 475–486 (2017).Article 

    Google Scholar 
    101.Kimura, M. On the probability of fixation of mutant genes in a population. Genetics 47, 713–719 (1962).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    102.Baguette, M. & Van Dyck, H. Landscape connectivity and animal behavior: functional grain as a key determinant for dispersal. Landsc. Ecol. 22, 1117–1129 (2007).Article 

    Google Scholar 
    103.Crow, J. F. & Aoki, K. Group selection for a polygenic behavioral trait: estimating the degree of population subdivision. Proc. Natl Acad. Sci. USA 81, 6073–6077 (1984).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    104.Lanner, R. Why do trees live so long? Ageing Res. Rev. 1, 653–671 (2002).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    105.Nabholz, B., Mauffrey, J.-F., Bazin, E., Galtier, N. & Glemin, S. Determination of mitochondrial genetic diversity in mammals. Genetics 178, 351–361 (2008).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    106.Lasne, C., Heerwaarden, B., Sgrò, C. M. & Connallon, T. Quantifying the relative contributions of the X chromosome, autosomes, and mitochondrial genome to local adaptation. Evolution 73, 262–277 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    107.Phillips, J. D., Gillis, D. J. & Hanner, R. H. Incomplete estimates of genetic diversity within species: implications for DNA barcoding. Ecol. Evol. 9, 2996–3010 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    108.Humphries, P. & Winemiller, K. O. Historical impacts on river fauna, shifting baselines, and challenges for restoration. BioScience 59, 673–684 (2009).Article 

    Google Scholar 
    109.Stoffel, M. A. et al. Demographic histories and genetic diversity across pinnipeds are shaped by human exploitation, ecology and life-history. Nat. Commun. 9, 4836 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    110.Collier-Robinson, L., Rayne, A., Rupene, M., Thoms, C. & Steeves, T. Embedding indigenous principles in genomic research of culturally significant species: a conservation genomics case study. N. Z. J. Ecol. 43, 3389 (2019).
    Google Scholar 
    111.Des Roches, S., Pendleton, L. H., Shapiro, B. & Palkovacs, E. P. Conserving intraspecific variation for nature’s contributions to people. Nat. Ecol. Evol. 5, 574–582 (2021).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    112.Gonzalez, A. et al. Estimating local biodiversity change: a critique of papers claiming no net loss of local diversity. Ecology 97, 1949–1960 (2016).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    113.Pope, L. C., Liggins, L., Keyse, J., Carvalho, S. B. & Riginos, C. Not the time or the place: the missing spatio-temporal link in publicly available genetic data. Mol. Ecol. 24, 3802–3809 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    114.Yilmaz, P. et al. Minimum information about a marker gene sequence (MIMARKS) and minimum information about any (x) sequence (MIxS) specifications. Nat. Biotechnol. 29, 415–420 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    115.Sibbett, B., Rieseberg, L. H. & Narum, S. The genomic observatories metadatabase. Mol. Ecol. Resour. 20, 1453–1454 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    116.Eichenberg, D. et al. Widespread decline in Central European plant diversity across six decades. Glob. Change Biol. 27, 1097–1110 (2020).Article 

    Google Scholar 
    117.Cornwell, W. K., Pearse, W. D., Dalrymple, R. L. & Zanne, A. E. What we (don’t) know about global plant diversity. Ecography 42, 1819–1831 (2019).Article 

    Google Scholar 
    118.Li, X. et al. Plant DNA barcoding: from gene to genome. Biol. Rev. 90, 157–166 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    119.Vasquez-Gross, H. A. et al. CartograTree: connecting tree genomes, phenotypes and environment. Mol. Ecol. Resour. 13, 528–537 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    120.Lawrence, E. R. et al. Geo-referenced population-specific microsatellite data across American continents, the MacroPopGen Database. Sci. Data 6, 14 (2019). This paper reports a compilation of georeferenced vertebrate microsatellite data, summary statistics and meta-data across the Americas for 897 species and 9,090 genetically distinct populations.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    121.Zellweger, F., De Frenne, P., Lenoir, J., Rocchini, D. & Coomes, D. Advances in microclimate ecology arising from remote sensing. Trends Ecol. Evol. 34, 327–341 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    122.Barber, P. H. et al. Advancing biodiversity research in developing countries: the need for changing paradigms. Bull. Mar. Sci. 90, 187–210 (2014).Article 

    Google Scholar 
    123.Bork, P. et al. Tara Oceans. Tara Oceans studies plankton at planetary scale. Introduction. Science 348, 873–873 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    124.Lotterhos, K. E. & Whitlock, M. C. The relative power of genome scans to detect local adaptation depends on sampling design and statistical method. Mol. Ecol. 24, 1031–1046 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    125.Hoban, S. et al. Genetic diversity targets and indicators in the CBD post-2020 Global Biodiversity Framework must be improved. Biol. Conserv. 248, 108654 (2020).Article 

    Google Scholar 
    126.Holmes, M. W. et al. Natural history collections as windows on evolutionary processes. Mol. Ecol. 25, 864–881 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    127.Boukhdoud, L. et al. First DNA sequence reference library for mammals and plants of the Eastern Mediterranean Region. Genome 64, 39–49 (2021).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    128.Colella, J. P. et al. The Open-Specimen movement. BioScience 71, 405–414 (2020).Article 

    Google Scholar 
    129.Wright, S. Correlation and causation. J. Agric. Res. 20, 557–585 (1921).
    Google Scholar 
    130.Fourtune, L. et al. Inferring causalities in landscape genetics: an extension of Wright’s causal modeling to distance matrices. Am. Nat. 191, 491–508 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    131.Paz-Vinas, I., Loot, G., Stevens, V. M. & Blanchet, S. Evolutionary processes driving spatial patterns of intraspecific genetic diversity in river ecosystems. Mol. Ecol. 24, 4586–4604 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    132.Beaumont, M. A., Zhang, W. & Balding, D. J. Approximate Bayesian computation in population genetics. Genetics 162, 2025–2035 (2002).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    133.Breiman, L. Random forests. Mach. Learn. 45, 5–32 (2001).Article 

    Google Scholar 
    134.Proença, V. et al. Global biodiversity monitoring: From data sources to Essential Biodiversity Variables. Biol. Conserv. 213, 256–263 (2017).Article 

    Google Scholar 
    135.Ve˅trovský, T. et al. A meta-analysis of global fungal distribution reveals climate-driven patterns. Nat. Commun. 10, 5142 (2019).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    136.Hanson, J. O. et al. Conservation planning for adaptive and neutral evolutionary processes. J. Appl. Ecol. 57, 2159–2169 (2020).Article 

    Google Scholar 
    137.Xuereb, A., D’Aloia, C. C., Andrello, M., Bernatchez, L. & Fortin, M. Incorporating putatively neutral and adaptive genomic data into marine conservation planning. Conserv. Biol. 35, 909–920 (2021).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    138.Carvalho, S. B., Torres, J., Tarroso, P. & Velo-Antón, G. Genes on the edge: a framework to detect genetic diversity imperiled by climate change. Glob. Change Biol. 25, 4034–4047 (2019).Article 

    Google Scholar 
    139.Adams, W. M. & Sandbrook, C. Conservation, evidence and policy. Oryx 47, 329–335 (2013).Article 

    Google Scholar 
    140.Laikre, L. et al. Post-2020 goals overlook genetic diversity. Science 367, 1083.2–1085 (2020).Article 
    CAS 

    Google Scholar 
    141.Thomson, A. I. et al. Charting a course for genetic diversity in the UN Decade of Ocean Science. Evol. Appl. 14, 1497–1518 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    142.Hoban, S. M. et al. Bringing genetic diversity to the forefront of conservation policy and management. Conserv. Genet. Resour. 5, 593–598 (2013).Article 

    Google Scholar 
    143.Carroll, S. R. et al. The CARE principles for indigenous data governance. Data Sci. J. 19, 43 (2020).Article 

    Google Scholar 
    144.Fargeot, L. et al. Patterns of epigenetic diversity in two sympatric fish species: genetic vs. environmental determinants. Genes 12, 107 (2021).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    145.Gaggiotti, O. E. et al. Diversity from genes to ecosystems: a unifying framework to study variation across biological metrics and scales. Evol. Appl. 11, 1176–1193 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    146.Waples, R. S., Antao, T. & Luikart, G. Effects of overlapping generations on linkage disequilibrium estimates of effective population size. Genetics 197, 769–780 (2014).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    147.Waples, R. S. & Yokota, M. Temporal estimates of effective population size in species with overlapping generations. Genetics 175, 219–233 (2007).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    148.Antao, T., Pérez-Figueroa, A. & Luikart, G. Early detection of population declines: high power of genetic monitoring using effective population size estimators. Evol. Appl. 4, 144–154 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    149.Cornuet, J. M. & Luikart, G. Description and power analysis of two tests for detecting recent population bottlenecks from allele frequency data. Genetics 144, 2001–2014 (1996).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    150.Phillips, J. D., Gwiazdowski, R. A., Ashlock, D. & Hanner, R. An exploration of sufficient sampling effort to describe intraspecific DNA barcode haplotype diversity: examples from the ray-finned fishes (Chordata: Actinopterygii). DNA Barcodes 3, 66–73 (2015).Article 

    Google Scholar 
    151.Tajima, F. The effect of change in population size on DNA polymorphism. Genetics 123, 597–601 (1989).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    152.Jordan, R., Breed, M. F., Prober, S. M., Miller, A. D. & Hoffmann, A. A. How well do revegetation plantings capture genetic diversity? Biol. Lett. 15, 20190460 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    153.Holderegger, R. & Di Giulio, M. The genetic effects of roads: a review of empirical evidence. Basic. Appl. Ecol. 11, 522–531 (2010).Article 

    Google Scholar 
    154.Hale, M. L., Burg, T. M. & Steeves, T. E. Sampling for microsatellite-based population genetic studies: 25 to 30 individuals per population is enough to accurately estimate allele frequencies. PLoS One 7, e45170 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    155.Jackson, T. M., Roegner, G. C. & O’Malley, K. G. Evidence for interannual variation in genetic structure of Dungeness crab (Cancer magister) along the California Current System. Mol. Ecol. 27, 352–368 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    156.Hoban, S. et al. Comparative evaluation of potential indicators and temporal sampling protocols for monitoring genetic erosion. Evol. Appl. 7, 984–998 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    157.Anderson, C. N. K., Ramakrishnan, U., Chan, Y. L. & Hadly, E. A. Serial SimCoal: a population genetics model for data from multiple populations and points in time. Bioinformatics 21, 1733–1734 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    158.Hortal, J. et al. Seven shortfalls that beset large-scale knowledge of biodiversity. Annu. Rev. Ecol. Evol. Syst. 46, 523–549 (2015).Article 

    Google Scholar 
    159.Elbrecht, V., Vamos, E. E., Steinke, D. & Leese, F. Estimating intraspecific genetic diversity from community DNA metabarcoding data. PeerJ 6, e4644 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    160.Shum, P. & Palumbi, S. R. Testing small-scale ecological gradients and intraspecific differentiation for hundreds of kelp forest species using haplotypes from metabarcoding. Mol. Ecol. https://doi.org/10.1111/mec.15851 (2021).Article 
    PubMed 

    Google Scholar 
    161.Yamahara, K. M. et al. In situ autonomous acquisition and preservation of marine environmental DNA using an autonomous underwater vehicle. Front. Mar. Sci. 6, 373 (2019).Article 

    Google Scholar 
    162.Breed, M. F. et al. Mating patterns and pollinator mobility are critical traits in forest fragmentation genetics. Heredity 115, 108–114 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    163.Hoban, S., Gaggiotti, O. & Bertorelle, G. Sample Planning Optimization Tool for conservation and population Genetics (SPOTG): a software for choosing the appropriate number of markers and samples. Methods Ecol. Evol. 4, 299–303 (2013).Article 

    Google Scholar 
    164.Peck, S. L. Simulation as experiment: a philosophical reassessment for biological modeling. Trends Ecol. Evol. 19, 530–534 (2004).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    165.Reid, B. N., Naro-Maciel, E., Hahn, A. T., FitzSimmons, N. N. & Gehara, M. Geography best explains global patterns of genetic diversity and postglacial co-expansion in marine turtles. Mol. Ecol. 28, 3358–3370 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    166.Kardos, M., Luikart, G. & Allendorf, F. W. Measuring individual inbreeding in the age of genomics: marker-based measures are better than pedigrees. Heredity 115, 63–72 (2015).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    167.Willing, E.-M., Dreyer, C. & van Oosterhout, C. Estimates of genetic differentiation measured by FST do not necessarily require large sample sizes when using many SNP markers. PLoS One 7, e42649 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    168.Shafer, A. B. A. et al. Bioinformatic processing of RAD-seq data dramatically impacts downstream population genetic inference. Methods Ecol. Evol. 8, 907–917 (2017).Article 

    Google Scholar 
    169.Cariou, M., Duret, L. & Charlat, S. How and how much does RAD-seq bias genetic diversity estimates? BMC Evol. Biol. 16, 240 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    170.De-Kayne, R. et al. Sequencing platform shifts provide opportunities but pose challenges for combining genomic data sets. Mol. Ecol. Resour. 21, 653–660 (2021).PubMed 
    Article 
    CAS 

    Google Scholar 
    171.Leigh, D. M., Lischer, H. E. L., Grossen, C. & Keller, L. F. Batch effects in a multiyear sequencing study: false biological trends due to changes in read lengths. Mol. Ecol. Resour. 18, 778–788 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    172.Linck, E. & Battey, C. J. Minor allele frequency thresholds strongly affect population structure inference with genomic data sets. Mol. Ecol. Resour. 19, 639–647 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    173.Benestan, L. M. et al. Conservation genomics of natural and managed populations: building a conceptual and practical framework. Mol. Ecol. 25, 2967–2977 (2016).PubMed 
    Article 

    Google Scholar 
    174.Feng, S. et al. Dense sampling of bird diversity increases power of comparative genomics. Nature 587, 252–257 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    175.Brandies, P., Peel, E., Hogg, C. J. & Belov, K. The value of reference genomes in the conservation of threatened species. Genes 10, 846 (2019).CAS 
    PubMed Central 
    Article 

    Google Scholar  More

  • in

    Quantitative mapping and spectroscopic characterization of particulate organic matter fractions in soil profiles with imaging VisNIR spectroscopy

    1.Smith, P. et al. The changing faces of soil organic matter research. Eur. J. Soil Sci. 69, 23–30. https://doi.org/10.1111/ejss.12500 (2018).Article 

    Google Scholar 
    2.Kögel-Knabner, I. The macromolecular organic composition of plant and microbial residues as inputs to soil organic matter. Soil Biol. Biochem. 34, 139–162 (2002).Article 

    Google Scholar 
    3.Schmidt, M. W. I. et al. Persistence of soil organic matter as an ecosystem property. Nature 478, 49–56. https://doi.org/10.1038/nature10386 (2011).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    4.Lehmann, J. & Kleber, M. The contentious nature of soil organic matter. Nature 528, 60–68. https://doi.org/10.1038/nature16069 (2015).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    5.Lehmann, J. et al. Persistence of soil organic carbon caused by functional complexity. Nat. Geosci. 13, 529–534. https://doi.org/10.1038/s41561-020-0612-3 (2020).ADS 
    CAS 
    Article 

    Google Scholar 
    6.Dong, L. et al. Effect of grazing exclusion and rotational grazing on labile soil organic carbon in north China. Eur. J. Soil Sci. https://doi.org/10.1111/ejss.12952 (2020).Article 

    Google Scholar 
    7.Leifeld, J. & Kogel-Knabner, I. Soil organic matter fractions as early indicators for carbon stock changes under different land-use?. Geoderma 124, 143–155 (2005).ADS 
    CAS 
    Article 

    Google Scholar 
    8.Poeplau, C. & Don, A. Sensitivity of soil organic carbon stocks and fractions to different land-use changes across Europe. Geoderma 192, 189–201. https://doi.org/10.1016/j.geoderma.2012.08.003 (2013).ADS 
    CAS 
    Article 

    Google Scholar 
    9.Besnard, E., Chenu, C., Balesdent, J., Puget, P. & Arrouays, D. Fate of particulate organic matter in soil aggregates during cultivation. Eur. J. Soil Sci. 47, 495–503 (1996).CAS 
    Article 

    Google Scholar 
    10.von Lützow, M. et al. Stabilization of organic matter in temperate soils: mechanisms and their relevance under different soil conditions—a review. Eur. J. Soil Sci. 57, 426–445. https://doi.org/10.1111/j.1365-2389.2006.00809.x (2006).CAS 
    Article 

    Google Scholar 
    11.Peng, X. H., Zhu, Q. H., Zhang, Z. B. & Hallett, P. D. Combined turnover of carbon and soil aggregates using rare earth oxides and isotopically labelled carbon as tracers. Soil Biol. Biochem. 109, 81–94. https://doi.org/10.1016/j.soilbio.2017.02.002 (2017).CAS 
    Article 

    Google Scholar 
    12.Dynarski, K. A., Bossio, D. A. & Scow, K. M. Dynamic stability of soil carbon: reassessing the “permanence” of soil carbon sequestration. Front. Environ. Sci. 8, 1. https://doi.org/10.3389/fenvs.2020.514701 (2020).Article 

    Google Scholar 
    13.Basile-Doelsch, I., Balesdent, J. & Pellerin, S. Reviews and syntheses: The mechanisms underlying carbon storage in soil. Biogeosci. Discuss. https://doi.org/10.5194/bg-2020-49 (2020).14.Poeplau, C. et al. Isolating organic carbon fractions with varying turnover rates in temperate agricultural soils—a comprehensive method comparison. Soil Biol. Biochem. 125, 10–26. https://doi.org/10.1016/j.soilbio.2018.06.025 (2018).CAS 
    Article 

    Google Scholar 
    15.Rumpel, C. & Kögel-Knabner, I. Deep soil organic matter-a key but poorly understood component of terrestrial C cycle. Plant Soil 338, 143–158. https://doi.org/10.1007/s11104-010-0391-5 (2011).CAS 
    Article 

    Google Scholar 
    16.Steffens, M., Kölbl, A., Schörk, E., Gschrey, B. & Kögel-Knabner, I. Distribution of soil organic matter between fractions and aggregate size classes in grazed semiarid steppe soil profiles. Plant Soil 338, 63–81. https://doi.org/10.1007/s11104-010-0594-9 (2011).CAS 
    Article 

    Google Scholar 
    17.Soriano-Disla, J. M., Janik, L. J., Rossel, R. A. V., Macdonald, L. M. & McLaughlin, M. J. The performance of visible, near-, and mid-infrared reflectance spectroscopy for prediction of soil physical, chemical, and biological properties. Appl. Spectrosc. Rev. 49, 139–186. https://doi.org/10.1080/05704928.2013.811081 (2014).ADS 
    CAS 
    Article 

    Google Scholar 
    18.Stenberg, B., Rossel, R. A. V., Mouazen, A. M. & Wetterlind, J. Visible and near infrared spectroscopy in soil science. Adv. Agron. 107(107), 163–215. https://doi.org/10.1016/s0065-2113(10)07005-7 (2010).CAS 
    Article 

    Google Scholar 
    19.Mouazen, A. M., Steffens, M. & Borisover, M. Reflectance and fluorescence spectroscopy in soil science-Current and future research and developments. Soil Tillage Res. 155, 448–449 (2016).Article 

    Google Scholar 
    20.Viscarra Rossel, R. A. & Bouma, J. Soil sensing: A new paradigm for agriculture. Agric. Syst. 148, 71–74. https://doi.org/10.1016/j.agsy.2016.07.001 (2016).Article 

    Google Scholar 
    21.Nocita, M. et al.. Soil spectroscopy: An alternative to wet chemistry for soil monitoring. Adv. Agron. 132, 139–159 (2015)Article 

    Google Scholar 
    22.Gholizadeh, A., Boruvka, L., Saberioon, M. & Vasat, R. Visible, near-infrared, and mid-infrared spectroscopy applications for soil assessment with emphasis on soil organic matter content and quality: state-of-the-art and key issues. Appl. Spectrosc. 67, 1349–1362. https://doi.org/10.1366/13-07288 (2013).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    23.Hermansen, C. et al. Complete soil texture is accurately predicted by visible near-infrared spectroscopy. Soil Sci. Soc. Am. J. 81, 758–769. https://doi.org/10.2136/sssaj2017.02.0066 (2017).ADS 
    CAS 
    Article 

    Google Scholar 
    24.Zimmermann, M., Leifeld, J. & Fuhrer, J. Quantifying soil organic carbon fractions by infrared-spectroscopy. Soil Biol. Biochem. 39, 224–231. https://doi.org/10.1016/j.soilbio.2006.07.010 (2007).CAS 
    Article 

    Google Scholar 
    25.Madhavan, D. B. et al. Rapid prediction of particulate, humus and resistant fractions of soil organic carbon in reforested lands using infrared spectroscopy. J. Environ. Manage. 193, 290–299. https://doi.org/10.1016/jjenvman.2017.02.013 (2017).CAS 
    Article 
    PubMed 

    Google Scholar 
    26.St. Luce, M. et al. Rapid determination of soil organic matter quality indicators using visible near infrared reflectance spectroscopy. Geoderma 232–234, 449–458. https://doi.org/10.1016/j.geoderma.2014.05.023 (2014).ADS 
    CAS 
    Article 

    Google Scholar 
    27.Terhoeven-Urselmans, T., Michel, K., Helfrich, M., Flessa, H. & Ludwig, B. Near-infrared spectroscopy can predict the composition of organic matter in soil and litter. J. Plant Nutr. Soil Sci. 169, 168–174. https://doi.org/10.1002/jpln.200521712 (2006).CAS 
    Article 

    Google Scholar 
    28.Margenot, A., O’Neill, T., Sommer, R. & Akella, V. Predicting soil permanganate oxidizable carbon (PDXC) by coupling DRIFT spectroscopy and artificial neural networks (ANN). Comput. Electron. Agric. https://doi.org/10.1016/j.compag.2019.105098 (2020).Article 

    Google Scholar 
    29.Fang, Q. et al. Visible and near-infrared reflectance spectroscopy for investigating soil mineralogy: a review. J. Spectrosc. https://doi.org/10.1155/2018/3168974 (2018).Article 

    Google Scholar 
    30.Shi, P., Castaldi, F., van Wesemael, B. & van Oost, K. Vis-NIR spectroscopic assessment of soil aggregate stability and aggregate size distribution in the Belgian Loam Belt. Geoderma https://doi.org/10.1016/j.geoderma.2019.113958 (2020).Article 

    Google Scholar 
    31.Canasveras, J. C., Barron, V., del Campillo, M. C., Torrent, J. & Gomez, J. A. Estimation of aggregate stability indices in Mediterranean soils by diffuse reflectance spectroscopy. Geoderma 158, 78–84. https://doi.org/10.1016/j.geoderma.2009.09.004 (2010).ADS 
    CAS 
    Article 

    Google Scholar 
    32.Hermansen, C. et al. Visible-near-infrared spectroscopy can predict the clay/organic carbon and mineral fines/organic carbon ratios. Soil Sci. Soc. Am. J. 80, 1486–1495. https://doi.org/10.2136/sssaj2016.05.0159 (2016).ADS 
    CAS 
    Article 

    Google Scholar 
    33.Jaconi, A., Don, A. & Freibauer, A. Prediction of soil organic carbon at the country scale: stratification strategies for near-infrared data. Eur. J. Soil Sci. 68, 919–929. https://doi.org/10.1111/ejss.12485 (2017).CAS 
    Article 

    Google Scholar 
    34.Jaconi, A., Vos, C. & Don, A. Near infrared spectroscopy as an easy and precise method to estimate soil texture. Geoderma 337, 906–913. https://doi.org/10.1016/j.geoderma.2018.10.038 (2019).ADS 
    Article 

    Google Scholar 
    35.Riedel, F., Denk, M., Muller, I., Barth, N. & Glasser, C. Prediction of soil parameters using the spectral range between 350 and 15,000 nm: A case study based on the Permanent Soil Monitoring Program in Saxony Germany. Geoderma 315, 188–198. https://doi.org/10.1016/j.geoderma.2017.11.027 (2018).ADS 
    CAS 
    Article 

    Google Scholar 
    36.Clairotte, M. et al. National calibration of soil organic carbon concentration using diffuse infrared reflectance spectroscopy. Geoderma 276, 41–52. https://doi.org/10.1016/j.geoderma.2016.04.021 (2016).ADS 
    CAS 
    Article 

    Google Scholar 
    37.Orgiazzi, A., Ballabio, C., Panagos, P., Jones, A. & Fernandez-Ugalde, O. LUCAS soil, the largest expandable soil dataset for Europe: a review. Eur. J. Soil Sci. 69, 140–153. https://doi.org/10.1111/ejss.12499 (2018).Article 

    Google Scholar 
    38.Stevens, A., Nocita, M., Toth, G., Montanarella, L. & van Wesemael, B. Prediction of soil organic carbon at the european scale by visible and near infrared reflectance spectroscopy. PLoS ONE 8, 1. https://doi.org/10.1371/journal.pone.0066409 (2013).CAS 
    Article 

    Google Scholar 
    39.Nocita, M. et al. Prediction of soil organic carbon content by diffuse reflectance spectroscopy using a local partial least square regression approach. Soil Biol. Biochem. 68, 337–347. https://doi.org/10.1016/j.soilbio.2013.10.022 (2014).CAS 
    Article 

    Google Scholar 
    40.Viscarra Rossel, R. A. & Hicks, W. S. Soil organic carbon and its fractions estimated by visible-near infrared transfer functions. Europ. J. Soil Sci. 66, 438–450. https://doi.org/10.1111/ejss.12237 (2015).CAS 
    Article 

    Google Scholar 
    41.Steffens, M. & Buddenbaum, H. Laboratory imaging spectroscopy of a stagnic Luvisol profile – High resolution soil characterisation, classification and mapping of elemental concentrations. Geoderma 195–196, 122–132 (2013).ADS 
    Article 

    Google Scholar 
    42.Steffens, M., Kohlpaintner, M. & Buddenbaum, H. Fine spatial resolution mapping of soil organic matter quality in a Histosol profile. Eur. J. Soil Sci. 65, 827–839. https://doi.org/10.1111/ejss.12182 (2014).CAS 
    Article 

    Google Scholar 
    43.Hobley, E., Steffens, M., Bauke, S. L. & Kogel-Knabner, I. Hotspots of soil organic carbon storage revealed by laboratory hyperspectral imaging. Sci. Rep. 8, 1. https://doi.org/10.1038/s41598-018-31776-w (2018).ADS 
    CAS 
    Article 

    Google Scholar 
    44.Lucas, M., Pihlap, E., Steffens, M., Vetterlein, D. & Kogel-Knabner, I. Combination of imaging infrared spectroscopy and x-ray computed microtomography for the investigation of bio- and physicochemical processes in structured soils. Front. Environ. Sci. 8, 1. https://doi.org/10.3389/fenvs.2020.00042 (2020).Article 

    Google Scholar 
    45.Mueller, C. W., Steffens, M. & Buddenbaum, H. Permafrost soil complexity evaluated by laboratory imaging Vis-NIR spectroscopy. Eur. J. Soil Sci. https://doi.org/10.1111/ejss.12927 (2019).Article 

    Google Scholar 
    46.Schreiner, S., Buddenbaum, H., Emmerling, C. & Steffens, M. VNIR/SWIR laboratory imaging spectroscopy for wall-to-wall mapping of elemental concentrations in soil cores. Photogrammetrie Fernerkundung Geoinformation https://doi.org/10.1127/pfg/2015/0279 (2015).Article 

    Google Scholar 
    47.Askari, M. S., O’Rourke, S. M. & Holden, N. M. A comparison of point and imaging visible-near infrared spectroscopy for determining soil organic carbon. J. Near Infrared Spectrosc. 26, 133–146. https://doi.org/10.1177/0967033518766668 (2018).ADS 
    CAS 
    Article 

    Google Scholar 
    48.O’Rourke, S. M. & Holden, N. M. Determination of soil organic matter and carbon fractions in forest top soils using spectral data acquired from visible-near infrared hyperspectral images. Soil Sci. Soc. Am. J. 76, 586–596. https://doi.org/10.2136/sssaj2011.0053 (2012).ADS 
    CAS 
    Article 

    Google Scholar 
    49.Buddenbaum, H. & Steffens, M. Laboratory imaging spectroscopy of soil profiles. J. Spectral Imag. 2, 1. https://doi.org/10.1255/jsi.2011.a2 (2011).Article 

    Google Scholar 
    50.Buddenbaum, H. & Steffens, M. Mapping the distribution of chemical properties in soil profiles using laboratory imaging spectroscopy SVM and PLS regression. EARSeL eProc. 11, 25–32 (2012).
    Google Scholar 
    51.Poeplau, C. et al. Stocks of organic carbon in German agricultural soils-Key results of the first comprehensive inventory. J. Plant Nutr. Soil Sci. 183, 665–681. https://doi.org/10.1002/jpln.202000113 (2020).CAS 
    Article 

    Google Scholar 
    52.Viscarra Rossel, R. A., Lobsey, C. R., Sharman, C., Flick, P. & McLachlan, G. Novel proximal sensing for monitoring soil organic C stocks and condition. Environ. Sci. Technol. 51, 5630–5641. https://doi.org/10.1021/acs.est.7b00889 (2017).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    53.IUSS Working Group WRB. World reference base for soil resources 2006. Vol. 103 (FAO, 2006).
    Google Scholar 
    54.Steffens, M., Kölbl, A., Totsche, K. U. & Kögel-Knabner, I. Grazing effects on soil chemical and physical properties in a semiarid steppe of Inner Mongolia (P.R. China). Geoderma 143, 63–72 (2008).ADS 
    CAS 
    Article 

    Google Scholar 
    55.Hoffmann, C. et al. Effects of grazing and climate variability on grassland ecosystem functions in Inner Mongolia: Synthesis of a 6-year grazing experiment. J. Arid Environ. 135, 50–63. https://doi.org/10.1016/j.jaridenv.2016.08.003 (2016).ADS 
    Article 

    Google Scholar 
    56.FAO. Guidelines for soil description. 4th edition edn, (FAO, 2006).
    Google Scholar 
    57.Lenhard, K., Baumgartner, A. & Schwarzmaier, T. Independent laboratory characterization of NEO HySpex imaging spectrometers VNIR-1600 and SWIR-320m-e. IEEE Trans. Geosci. Remote Sens. 53, 1828–1841. https://doi.org/10.1109/TGRS.2014.2349737 (2015).ADS 
    Article 

    Google Scholar 
    58.Peddle, D. R., White, H. P., Soffer, R. J., Miller, J. R. & LeDrew, E. F. Reflectance processing of remote sensing spectroradiometer data. Comput. Geosci. 27, 203–213 (2001).ADS 
    Article 

    Google Scholar 
    59.Rogass, C. et al. Translational imaging spectroscopy for proximal sensing. Sensors 17, 1857 (2017).Article 

    Google Scholar 
    60.Steffens, M., Kölbl, A. & Kögel-Knabner, I. Alteration of soil organic matter pools and aggregation in semi-arid steppe topsoils as driven by organic matter input. Eur. J. Soil Sci. 60, 198–212. https://doi.org/10.1111/j.1365-2389.2008.01104.x (2009).CAS 
    Article 

    Google Scholar 
    61.Golchin, A., Oades, J. M., Skjemstad, J. O. & Clarke, P. Soil-structure and carbon cycling. Aust. J. Soil Res. 32, 1043–1068 (1994).Article 

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

    Google Scholar 
    63.Schmidt, M. W. I., Rumpel, C. & Kögel-Knabner, I. Evaluation of an ultrasonic dispersion procedure to isolate primary organomineral complexes from soils. Eur. J. Soil Sci. 50, 87–94 (1999).Article 

    Google Scholar 
    64.Steffens, M. et al. Spatial variability of topsoils and vegetation in a grazed steppe ecosystem in Inner Mongolia (PR China). J. Plant Nutr. Soil Sci. 172, 78–90. https://doi.org/10.1002/jpln.200700309 (2009).CAS 
    Article 

    Google Scholar 
    65.Six, J., Gregorich, E. & Koegel-Knabner, I. Landmark Papers: No. 1. Tisdall, J. M. & Oades, J. M. 1982. Organic matter and water-stable aggregates in soils. Journal of Soil Science, 33, 141–163 Commentary on the impact of the impact of Tisdall & Oades (1982): by J. Six, E. G. Gregorich & I. Kogel-Knabner. European Journal of Soil Science 63, 3–7 (2012).66.Six, J., Bossuyt, H., Degryze, S. & Denef, K. A history of research on the link between (micro)aggregates, soil biota, and soil organic matter dynamics. Soil Tillage Res. 79, 7–31 (2004).Article 

    Google Scholar 
    67.Wiesmeier, M. et al. Aggregate stability and physical protection of soil organic carbon in semi-arid steppe soils. Eur. J. Soil Sci. 63, 22–31. https://doi.org/10.1111/j.1365-2389.2011.01418.x (2012).CAS 
    Article 

    Google Scholar 
    68.McSherry, M. E. & Ritchie, M. E. Effects of grazing on grassland soil carbon: a global review. Glob. Change Biol. 19, 1347–1357. https://doi.org/10.1111/gcb.12144 (2013).ADS 
    Article 

    Google Scholar 
    69.Viscarra Rossel, R. A. & Behrens, T. Using data mining to model and interpret soil diffuse reflectance spectra. Geoderma 158, 46–54. https://doi.org/10.1016/j.geoderma.2009.12.025 (2010).ADS 
    CAS 
    Article 

    Google Scholar 
    70.Ben-Dor, E., Inbar, Y. & Chen, Y. The reflectance spectra of organic matter in the visible near-infrared and short wave infrared region (400–2500 nm) during a controlled decomposition process. Remote Sens. Environ. 61, 1–15 (1997).ADS 
    Article 

    Google Scholar 
    71.Delegido, J., Verrelst, J., Rivera, J. P., Ruiz-Verdu, A. & Moreno, J. Brown and green LAI mapping through spectral indices. Int. J. Appl. Earth Obs. Geoinf. 35, 350–358. https://doi.org/10.1016/j.jag.2014.10.001 (2015).ADS 
    Article 

    Google Scholar 
    72.Viscarra Rossel, R. A., McGlynn, R. N. & McBratney, A. B. Determing the composition of mineral-organic mixes using UV-vis-NIR diffuse reflectance spectroscopy. Geoderma 137, 70–82. https://doi.org/10.1016/j.geoderma.2006.07.004 (2006).ADS 
    CAS 
    Article 

    Google Scholar 
    73.Ben-Dor, E. et al. Imaging spectrometry for soil applications. Adv. Agronomy 97, 321. https://doi.org/10.1016/s0065-2113(07)00008-9 (2008).CAS 
    Article 

    Google Scholar  More

  • in

    Stochasticity in host-parasitoid models informs mechanisms regulating population dynamics

    1.Benincà, E., Ballantine, B., Ellner, S.P. & Huisman, J. Species fluctuations sustained by a cyclic succession at the edge of chaos. Proc. Natl. Acad. Sci. 112, 6389–6394 (2015).2.Lande, R. et al. Stochastic Population Dynamics in Ecology and Conservation (Oxford University Press, 2003).3.Bonsall, M. B. & Hastings, A. Demographic and environmental stochasticity in predator-prey metapopulation dynamics. J. Anim. Ecol. 73, 1043–1055 (2004).Article 

    Google Scholar 
    4.Nisbet, R. M. & Gurney, W. Modelling Fluctuating Populations: reprint of first Edition (1982) (Blackburn Press, 2003).5.Hening, A. & Nguyen, D. H. Stochastic Lotka–Volterra food chains. J. Math. Biol. 77(1), 135–163 (2018).MathSciNet 
    Article 

    Google Scholar 
    6.Khasminskii, R. et al. Long term behavior of solutions of the Lotka–Volterra system under small random perturbations. Ann. Appl. Probab. 11(3), 952–963 (2001).MathSciNet 
    Article 

    Google Scholar 
    7.Huang, W., Hauert, C. & Traulsen, A. Stochastic game dynamics under demographic fluctuations. Proc. Natl. Acad. Sci., 112(29), 9064–9069 (2015).8.Suvinthra, M. & Balachandran, K. Large deviations for the stochastic predator-prey model with nonlinear functional response. J. Appl. Probab. 54(2), 507 (2017).MathSciNet 
    Article 

    Google Scholar 
    9.Zou, X. & Wang, K. Optimal harvesting for a stochastic Lotka–Volterra predator-prey system with jumps and nonselective harvesting hypothesis. Optim. Control Appl. Methods 37(4), 641–662 (2016).MathSciNet 
    Article 

    Google Scholar 
    10.Larsen, A. E. Modeling multiple nonconsumptive effects in simple food webs: a modified Lotka–Volterra approach. Behav. Ecol. 23(5), 1115–1125 (2012).Article 

    Google Scholar 
    11.Singh, A. Stochastic dynamics of consumer-resource interactions. bioRxiv (2021).12.Bashkirtseva, I., Ryashko, L. & Tsvetkov, I. Analysis of stochastic phenomena in ricker-type population model with delay. In AIP Conference Proceedings, vol. 1895, p. 050003 (2017).13.Halley, J. M. & Iwasa, Y. Extinction rate of a population under both demographic and environmental stochasticity. Theor. Popul. Biol. 53, 1–15 (1998).CAS 
    Article 

    Google Scholar 
    14.Hassell, M. P. (Oxford University Press, 2000).15.Gurney, W. S. C. & Nisbet, R. M. Ecological Dynamics (Oxford University Press, 1998).16.Murdoch, W. W., Briggs, C. J. & Nisbet, R. M. Consumer-Resouse Dynamics (Princeton University Press, 2003).17.Kakehashi, N., Suzuki, Y. & Iwasa, Y. Niche overlap of parasitoids in host-parasitoid systems: its consequence to single versus multiple introduction controversy in biological control. J. Appl. Ecol. 21, 115–131 (1984).Article 

    Google Scholar 
    18.May, R. M. & Hassell, M. P. The dynamics of multiparasitoid-host interactions. Am. Nat. 117(3), 234–261 (1981).MathSciNet 
    Article 

    Google Scholar 
    19.Hackett-Jones, E., Cobbold, C. & White, A. Coexistence of multiple parasitoids on a single host due to differences in parasitoid phenology. Theor. Ecol. 2(1), 19–31 (2009).Article 

    Google Scholar 
    20.van Velzen, E., Pérez-Vila, S. & Etienne, R. S. The role of within-host competition for coexistence in multiparasitoid-host systems. Am. Nat. 187(1), 48–59 (2016).Article 

    Google Scholar 
    21.Nicholson, A. & Bailey, V. A. The balance of animal populations. Part 1. Proc. Zool. Soc. Lond. 3, 551–598 (1935).Article 

    Google Scholar 
    22.Singh, A., Murdoch, W. W. & Nisbet, R. M. Skewed attacks, stability, and host suppression. Ecology 90(6), 1679–1686 (2009).Article 

    Google Scholar 
    23.Bešo, E., Kalabušić, S., Mujić, N. & Pilav, E. Stability of a certain class of a host-parasitoid models with a spatial refuge effect. J. Biol. Dyn. 14(1), 1–31 (2020).MathSciNet 
    Article 

    Google Scholar 
    24.Taylor, A. D. Heterogeneity in host-parasitoid interactions: ‘aggregation of risk’ and the (cv^2 >1) rule. Trends Ecol. Evolu. 8, 400–405 (1993).25.Hassell, M. P., May, R. M., Pacala, S. W. & Chesson, P. L. The persistence of host-parasitoid associations in patchy environments. I. A general criterion. Am. Nat. 138, 568–583 (1991).Article 

    Google Scholar 
    26.Pacala, S. W. & Hassell, M. P. The persistence of host- parasitoid associations in patchy environments. II. Evaluation of field data. Am. Nat. 138, 584–605 (1991).Article 

    Google Scholar 
    27.Bernstein, C. Density dependence and the stability of host-parasitoid systems. Oikos 47, 176–180 (1986).Article 

    Google Scholar 
    28.Free, C., Beddington, J. & Lawton, J. On the inadequacy of simple models of mutual interference for parasitism and predation. J. Anim. Ecol. 46, 543–554 (1977).Article 

    Google Scholar 
    29.Rogers, D. & Hassell, M. General models for insect parasite and predator searching behaviour: interference. J. Anim. Ecol. 43, 239–253 (1974).Article 

    Google Scholar 
    30.Reeve, J. D., Cronin, J. T. & Strong, D. R. Parasitoid aggregation and the stabilization of a salt marsh host- parasitoid system. Ecology 75, 288–295 (1994).Article 

    Google Scholar 
    31.Rohani, P., Godfray, H. C. J. & Hassell, M. P. Aggregation and the dynamics of host-parasitoid systems: A discrete-generation model with within-generation redistribution. Am. Nat. 144(3), 491–509 (1994).Article 

    Google Scholar 
    32.May, R. M. Host-parasitoid systems in patchy environments: A phenomenological model. J. Anim. Ecol. 47, 833–844 (1978).Article 

    Google Scholar 
    33.Singh, A. & Nisbet, R. M. Semi-discrete host-parasitoid models. J. Theor. Biol. 247(4), 733–742 (2007).ADS 
    MathSciNet 
    Article 

    Google Scholar 
    34.Singh, A. Population dynamics of multi-host communities attacked by a common parasitoid, bioRxiv (2021).35.Singh, A. & Emerick, B. Hybrid systems framework for modeling host-parasitoid population dynamics. In 2020 59th IEEE Conference on Decision and Control (CDC), 4628–4633 (2020).36.Lane, S. D., St, C. M. Mary, & Getz, W. M. Coexistence of attack-limited parasitoids sequentially exploiting the same resource and its implications for biological control. Ann. Zool. Fenn. 43, 17–34 (2006).
    Google Scholar 
    37.Pedersen, B. S. & Mills, N. J. Single vs. multiple introduction in biological control: the roles of parasitoid efficiency, antagonism and niche overlap. J. Appl. Ecol. 41(5), 973–984 (2004).Article 

    Google Scholar 
    38.Abram, P. K., Brodeur, J., Burte, V. & Boivin, G. Parasitoid-induced host egg abortion; an underappreciated component of biological control services provided by egg parasitoids. Biol. Control 98, 52–60 (2016).Article 

    Google Scholar 
    39.Jervis, M. A., Hawkin, B. A. & Kidd, N. A. C. The usefulness of destructive host-feeding parasitoids in classical biological control: Theory and observation conflict. Ecol. Entomol. 21(1), 41–46 (1996).Article 

    Google Scholar 
    40.Okuyama, T. Density-dependent distribution of parasitism risk among underground hosts. Bull. Entomol. Res. 109(4), 528–533 (2019).CAS 
    Article 

    Google Scholar 
    41.Cobbold, C. A., Roland, J. & Lewis, M. A. The impact of parasitoid emergence time on host-parastioid population dynamics. Theor. Popul. Biol. 75(2), 201–215 (2009).Article 

    Google Scholar 
    42.Liere, H., Jackson, D. & Vandermeer, J. Ecological complexity in a coffee agroecosystem: Spatial heterogeneity, popoulation persistence and biological control. PLoS One 7(9), e45508 (2012).43.Zoroa, N., Lesigne, E., Fernandez-Saez, M.J., Zoroa, P. & Casas, J. The coupon collector urn model with unequal probabilities in ecology and evolution, J. R. Soc. Interface 14, 20160643 (2017).44.Singh, A. & Emerick, B. Generalized stability conditions for host-parasitoid population dynamics: Implications for biological control. Ecol. Model. 456, 109656 (2021).45.Ledder, G. Mathematics for the Life Sciences: Calculus, Modeling, Probability, and Dynamical Systems (Springer Science & Business Media, 2013).46.Elaydi, S. An Introduction to Difference Equations (Springer, 1996).47.Gajic, Z. & Qureshi, M. T. J. Lyapunov matrix equation in system stability and control. (Courier Corporation, 2008).48.Singh, A. & Nisbet, R. M. Variation in risk in single-species discrete-time models. Math. Biosci. Eng. 5, 859–875 (2008).MathSciNet 
    Article 

    Google Scholar 
    49.Emerick, B. K. & Singh, A. The effects of host-feeding on stability of discrete-time host-parasitoid population dynamic models. Math. Biosci. 272, 54–63 (2016).MathSciNet 
    Article 

    Google Scholar 
    50.Pachepsky, E., Nisbet, R. M. & Murdoch, W. W. Between discrete and continuous: Consumer-resource dynamics with synchronized reproduction. Ecology 89(1), 280–288 (2007).Article 

    Google Scholar 
    51.Emerick, B. K., Singh, A & Chhetri, S. R. Global redistribution and local migration in semi-discrete host-parasitoid population dynamic models. Math. Biosci. 327, 108409 (2020).52.Rogers, D. J. Random searching and incest population models. J. Anim. Ecol. 41, 369–383 (1972).Article 

    Google Scholar 
    53.Hassell, M. P. & Comins, H. N. Sigmoid functional responses and population stability. Theor. Popul. Biol. 14, 62–66 (1978).CAS 
    Article 

    Google Scholar 
    54.Fernández-arhex, V. & Corley, J. C. The functional response of parasitoids and its implications for biological control. Biocontrol Sci. Technol. 13(4), 403–413 (2003).Article 

    Google Scholar 
    55.Okuyama, T. Dilution effects enhance variation in parasitism risk among hosts and stabilize host-parasitoid population dynamics. Ecol. Model. 441, 109425 (2021). More

  • in

    Species diversity and food web structure jointly shape natural biological control in agricultural landscapes

    1.van der Plas, F. Biodiversity and ecosystem functioning in naturally assembled communities. Biol. Rev. 94, 1220–1245 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    2.Lefcheck, J. S. et al. Biodiversity enhances ecosystem multifunctionality across trophic levels and habitats. Nat. Commun. 6, 6936 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    3.Fanin, N. et al. Consistent effects of biodiversity loss on multifunctionality across contrasting ecosystems. Nat. Ecol. Evol. 2, 269–278 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    4.Isbell, F. et al. Linking the influence and dependence of people on biodiversity across scales. Nature 546, 65–72 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    5.Sánchez-Bayo, F. & Wyckhuys, K. A. G. Worldwide decline of the entomofauna: A review of its drivers. Biol. Conserv. 232, 8–27 (2019).Article 

    Google Scholar 
    6.IPBES. Summary for policymakers of the global assessment report on biodiversity and ecosystem services of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services. (IPBES secretariat, Bonn, Germany, 2019).7.Smith, H. F. & Sullivan, C. A. Ecosystem services within agricultural landscapes—farmers’ perceptions. Ecol. Econ. 98, 72–80 (2014).Article 

    Google Scholar 
    8.Barnes, A. D. et al. Biodiversity enhances the multitrophic control of arthropod herbivory. Sci. Adv. 6, eabb6603 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    9.Dainese, M. et al. A global synthesis reveals biodiversity-mediated benefits for crop production. Sci. Adv. 5, eaax0121 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    10.Costanza, R. et al. The value of the world’s ecosystem services and natural capital. Nature 387, 253–260 (1997).CAS 
    Article 

    Google Scholar 
    11.Oliver, T. H. et al. Declining resilience of ecosystem functions under biodiversity loss. Nat. Commun. 6, 10122 (2015).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    12.Naranjo, S. E., Ellsworth, P. C. & Frisvold, G. B. Economic value of biological control in integrated pest management of managed plant systems. Annu. Rev. Entomol. 60, 621–645 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    13.Frishkoff, L. O. et al. Loss of avian phylogenetic diversity in neotropical agricultural systems. Science 345, 1343–1346 (2014).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    14.Mendenhall, C. D., Karp, D. S., Meyer, C. F. J., Hadly, E. A. & Daily, G. C. Predicting biodiversity change and averting collapse in agricultural landscapes. Nature 509, 213–217 (2014).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    15.Karp, D. S. et al. Crop pests and predators exhibit inconsistent responses to surrounding landscape composition. Proc. Natl Acad. Sci. USA 115, E7863–E7870 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    16.Tamburini, G. et al. Agricultural diversification promotes multiple ecosystem services without compromising yield. Sci. Adv. 6, eaba1715 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    17.Redlich, S., Martin, E. A. & Steffan-Dewenter, I. Landscape-level crop diversity benefits biological pest control. J. Appl. Ecol. 55, 2419–2428 (2018).Article 

    Google Scholar 
    18.Muneret, L. et al. Evidence that organic farming promotes pest control. Nat. Sustain. 1, 361–368 (2018).Article 

    Google Scholar 
    19.Roubos, C. R., Rodriguez-Saona, C. & Isaacs, R. Mitigating the effects of insecticides on arthropod biological control at field and landscape scales. Biol. Control 75, 28–38 (2014).CAS 
    Article 

    Google Scholar 
    20.Roschewitz, I., Hucker, M., Tscharntke, T. & Thies, C. The influence of landscape context and farming practices on parasitism of cereal aphids. Agric. Ecosyst. Environ. 108, 218–227 (2005).Article 

    Google Scholar 
    21.Frago, E., Pujadevillar, J., Guara, M. & Selfa, J. Hyperparasitism and seasonal patterns of parasitism as potential causes of low top-down control in Euproctis chrysorrhoea L. (Lymantriidae). Biol. Control 60, 123–131 (2012).Article 

    Google Scholar 
    22.Rosenheim, J. A., Kaya, H. K., Ehler, L. E., Marois, J. J. & Jaffee, B. A. Intraguild predation among biological-control agents: theory and evidence. Biol. Control 5, 303–335 (1995).Article 

    Google Scholar 
    23.Brobyn, P. J., Clark, S. J. & Wilding, N. The effect of fungus infection of Metopolophium dirhodum [Hom.: Aphididae] on the oviposition behaviour of the aphid parasitoid Aphidius rhopalosiphi [Hym.: Aphidiidae]. Entomophaga 33, 333–338 (1988).Article 

    Google Scholar 
    24.Tscharntke, T. et al. Conservation biological control and enemy diversity on a landscape scale. Biol. Control 43, 294–309 (2007).Article 

    Google Scholar 
    25.Rand, T. A., van Veen, F. J. F. & Tscharntke, T. Landscape complexity differentially benefits generalized fourth, over specialized third, trophic level natural enemies. Ecography 35, 97–104 (2012).Article 

    Google Scholar 
    26.Zhao, Z. H., Hui, C., He, D. H. & Li, B. L. Effects of agricultural intensification on ability of natural enemies to control aphids. Sci. Rep. 5, 8024 (2015).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    27.Vollhardt, I. M. G., Tscharntke, T., Wäckers, F. L., Bianchi, F. J. J. A. & Thies, C. Diversity of cereal aphid parasitoids in simple and complex landscapes. Agric. Ecosyst. Environ. 126, 289–292 (2008).Article 

    Google Scholar 
    28.Tomanović, Z. et al. Regional tritrophic relationship patterns of five aphid parasitoid species (Hymenoptera: Braconidae: Aphidiinae) in agroecosystem-dominated landscapes of southeastern Europe. J. Econ. Entomol. 102, 836–854 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    29.Kaartinen, R. & Roslin, T. Shrinking by numbers: landscape context affects the species composition but not the quantitative structure of local food webs. J. Anim. Ecol. 80, 622–631 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    30.Wang, S. & Brose, U. Biodiversity and ecosystem functioning in food webs: the vertical diversity hypothesis. Ecol. Lett. 21, 9–20 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    31.Garzke, J., Connor, S. J., Sommer, U. & O’Connor, M. I. Trophic interactions modify the temperature dependence of community biomass and ecosystem function. PLoS Biol. 17, e2006806 (2019).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    32.Pocock, M. J. O. et al. The visualisation of ecological networks, and their use as a tool for engagement, advocacy and management. Adv. Ecol. Res. 54, 41–85 (2016).Article 

    Google Scholar 
    33.Bersier, L.-F., Banašek-Richter, C. & Cattin, M.-F. Quantitative descriptors of food-web matrices. Ecology 83, 2394–2407 (2002).Article 

    Google Scholar 
    34.Tylianakis, J. M., Laliberté, E., Nielsen, A. & Bascompte, J. Conservation of species interaction networks. Biol. Conserv. 143, 2270–2279 (2010).Article 

    Google Scholar 
    35.Gilbert, A. J. Connectance indicates the robustness of food webs when subjected to species loss. Ecol. Indic. 9, 72–80 (2009).Article 

    Google Scholar 
    36.Williams, R. J. & Martinez, N. D. Simple rules yield complex food webs. Nature 404, 180–183 (2000).CAS 
    PubMed 
    Article 

    Google Scholar 
    37.Galiana, N., Hawkins, B. A. & Montoya, J. M. The geographical variation of network structure is scale dependent: understanding the biotic specialization of host–parasitoid networks. Ecography 42, 1175–1187 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    38.Banašek-Richter, C., Cattin, M.-F. & Bersier, L.-F. Sampling effects and the robustness of quantitative and qualitative food-web descriptors. J. Theor. Biol. 226, 23–32 (2004).PubMed 
    Article 

    Google Scholar 
    39.Varennes, Y. D., Boyer, S. & Wratten, S. D. Un-nesting DNA Russian dolls—the potential for constructing food webs using residual DNA in empty aphid mummies. Mol. Ecol. 23, 3925–3933 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    40.Zhu, Y. L. et al. A molecular detection approach for a cotton aphid-parasitoid complex in northern China. Sci. Rep. 9, 15836 (2019).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    41.Staniczenko, P. P. A. et al. Predicting the effect of habitat modification on networks of interacting species. Nat. Commun. 8, 792 (2018).Article 
    CAS 

    Google Scholar 
    42.Thies, C. & Tscharntke, T. In Biocontrol-Based Integrated Management of Oilseed Rape Pests (ed. Williams, I.H.). (Springer Netherlands, 2010).43.Tylianakis, J. M., Tscharntke, T. & Lewis, O. T. Habitat modification alters the structure of tropical host-parasitoid food webs. Nature 445, 202–205 (2007).CAS 
    Article 

    Google Scholar 
    44.Grass, I., Jauker, B., Steffandewenter, I., Tscharntke, T. & Jauker, F. Past and potential future effects of habitat fragmentation on structure and stability of plant-pollinator and host-parasitoid networks. Nat. Ecol. Evol. 2, 1408–1417 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    45.Gagic, V. et al. Food web structure and biocontrol in a four-trophic level system across a landscape complexity gradient. Proc. Roy. Soc. B. 278, 2946–2953 (2011).Article 

    Google Scholar 
    46.Lundgren, J. G. & Fausti, S. W. Trading biodiversity for pest problems. Sci. Adv. 1, e1500558 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    47.Zhou, K. et al. Effects of land use and insecticides on natural enemies of aphids in cotton: first evidence from smallholder agriculture in the North China Plain. Agric. Ecosyst. Environ. 183, 176–184 (2014).Article 

    Google Scholar 
    48.Zhang, Z. Q. The natural enemies of Aphis gossypii Glover (Hom., Aphididae) in China. J. Appl. Entomol. 114, 251–262 (2009).Article 

    Google Scholar 
    49.Gagic, V. et al. Agricultural intensification and cereal aphid–parasitoid–hyperparasitoid food webs: network complexity, temporal variability and parasitism rates. Oecologia 170, 1099–1109 (2012).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    50.Vollhardt, I. M. G. et al. Influence of plant fertilisation on cereal aphid-primary parasitoid-secondary parasitoid networks in simple and complex landscapes. Agric. Ecosyst. Environ. 281, 47–55 (2019).CAS 
    Article 

    Google Scholar 
    51.Sullivan, D. J. & Völkl, W. Hyperparasitism: multitrophic ecology and behavior. Annu. Rev. Entomol. 44, 291–315 (1999).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    52.Dainese, M., Montecchiari, S., Sitzia, T., Sigura, M. & Marini, L. High cover of hedgerows in the landscape supports multiple ecosystem services in Mediterranean cereal fields. J. Appl. Ecol. 54, 380–388 (2016).Article 

    Google Scholar 
    53.Landis, D. A., Wratten, S. D. & Gurr, G. M. Habitat management to conserve natural enemies of arthropod pests in agriculture. Annu. Rev. Entomol. 45, 175–201 (2000).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    54.Thies, C., Roschewitz, I. & Tscharntke, T. The landscape context of cereal aphid-parasitoid interactions. Proc. Roy. Soc. B. 272, 203–210 (2005).Article 

    Google Scholar 
    55.Plećaš, M. et al. Landscape composition and configuration influence cereal aphid–parasitoid–hyperparasitoid interactions and biological control differentially across years. Agric. Ecosyst. Environ. 183, 1–10 (2014).Article 

    Google Scholar 
    56.Sirami, C. et al. Increasing crop heterogeneity enhances multitrophic diversity across agricultural regions. Proc. Natl Acad. Sci. USA 116, 16442–16447 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    57.Lichtenberg, E. M. et al. A global synthesis of the effects of diversified farming systems on arthropod diversity within fields and across agricultural landscapes. Glob. Change Biol. 23, 4946–4957 (2017).Article 

    Google Scholar 
    58.Osorio, S., Arnan, X., Bassols, E., Vicens, N. & Bosch, J. Local and landscape effects in a host-parasitoid interaction network along a forest-cropland gradient. Ecol. Appl. 25, 1869–1879 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    59.Dunne, J., Williams, R. & Martinez, N. Network topology and biodiversity loss in food webs: robustness increases with connectance. Ecol. Lett. 5, 558–567 (2002).Article 

    Google Scholar 
    60.Montoya, J. M., Rodríguez, M. A. & Hawkins, B. A. Food web complexity and higher-level ecosystem services. Ecol. Lett. 6, 587–593 (2003).Article 

    Google Scholar 
    61.Hawkins, B. A. Parasitoid-host food webs and donor control. Oikos 65, 159–162 (1992).Article 

    Google Scholar 
    62.Yeakel, J. D. et al. Diverse interactions and ecosystem engineering can stabilize community assembly. Nat. Commun. 11, 3307 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    63.Poisot, T., Mouquet, N. & Gravel, D. Trophic complementarity drives the biodiversity-ecosystem functioning relationship in food webs. Ecol. Lett. 16, 853–861 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    64.White, L., O’Connor, N. E., Yang, Q., Emmerson, M. C. & Donohue, I. Individual species provide multifaceted contributions to the stability of ecosystems. Nat. Ecol. Evol. 4, 1594–1601 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    65.Ho, H.-C., Tylianakis, J. M. & Pawar, S. Behaviour moderates the impacts of food-web structure on species coexistence. Ecol. Lett. 24, 298–309 (2021).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    66.Holland, J. M. et al. Agri-environment scheme enhancing ecosystem services: A demonstration of improved biological control in cereal crops. Agric. Ecosyst. Environ. 155, 147–152 (2012).Article 

    Google Scholar 
    67.Batary, P., Dicks, L., Kleijn, D. & Sutherland, W. The role of agri-environment schemes in conservation and environmental management: European Agri-Environment Schemes. Conserv. Biol. 29, 1006–1016 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    68.McGarigal, K., Cushman, S., Maile, N. & Ene, E. FRAGSTATS v4: Spatial Pattern Analysis Program for Categorical and Continuous Maps. Computer software program produced by the authors at the University of Massachusetts, Amherst. http://www.umass.edu/landeco/research/fragstats/fragstats.html (2012).69.Liu, B. et al. Secondary crops and non-crop habitats within landscapes enhance the abundance and diversity of generalist predators. Agric. Ecosyst. Environ. 258, 30–39 (2018).Article 

    Google Scholar 
    70.Lu, Y. H., Qi, F. J. & Zhang, Y. J. Integrated Management of Diseases and Insect Pests in Cotton (Golden Shield Press, Beijing 2010).71.Shannon, C. E., Weaver, W., Blahut, R. E. & Hajek, B. The Mathematical Theory of Communications (University of Illinois Press, Urbana, 1949).72.Kembel, S. W. et al. Picante: R tools for integrating phylogenies and ecology. Bioinformatics 26, 1463–1464 (2010).CAS 
    Article 

    Google Scholar 
    73.R Development Core Team. R: A language and environment for statistical computing, Version 4.0.2. R Foundation for Statistical Computing, Vienna, Austria. http://www.R-project.org (2020).74.Dormann, C. F., Fründ, J. & Gruber, B. Package ‘bipartite’: Visualising bipartite networks and calculating some (ecological) indices. (2019).75.Huang, H. Y., Zhou, L., Chen, J. & Wei, T. Y. ggcor: Extended tools for correlation analysis and visualization. R package version 0.9.7. (2020).76.Oksanen, J. et al. vegan: community ecology package. R. package version 2, 5–6 (2020).
    Google Scholar 
    77.Kassambara, A. & Fabian, M. factoextra: Extract and Visualize the Results of Multivariate Data analyses. R package version 1.0.7. (2020).78.Akaike, H. An information criterion (AIC). Math. Sci. 14, 5–9 (1976).
    Google Scholar 
    79.Burnham, K. P. & Anderson, D. R. Multimodel inference understanding AIC and BIC in model selection. Sociol. Method. Res. 33, 261–304 (2004).Article 

    Google Scholar 
    80.Cardinale, B. J. et al. Effects of biodiversity on the functioning of trophic groups and ecosystems. Nature 443, 989–992 (2006).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    81.Fox, J. & Weisberg, S. An R Companion to Applied Regression, Third Edition. (Thousand Oaks CA: Sage., 2011).82.Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).Article 

    Google Scholar 
    83.Bartoń, K. MuMIn: Multi-Model Inference. R package version 1.43.17. (2020).84.Thompson, R. M. et al. Food webs: reconciling the structure and function of biodiversity. Trends Ecol. Evol. 27, 689–697 (2012).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    85.Tylianakis, J. M. & Binzer, A. Effects of global environmental changes on parasitoid–host food webs and biological control. Biol. Control 75, 77–86 (2014).Article 

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

    Google Scholar 
    87.Shipley, B. The AIC model selection method applied to path analytic models compared using a d-separation test. Ecology 94, 560–564 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    88.Yang, F. et al. The data for “Species diversity and food web structure jointly shape natural biological control in agricultural landscapes”. Dryad, Dataset https://doi.org/10.5061/dryad.pc866t1kz (2021).Article 

    Google Scholar  More

  • in

    Predator cue-induced plasticity of morphology and behavior in planthoppers facilitate the survival from predation

    To defend against predators, insects often modify their morphology, flexibly, to enhance survival and reproductive advantages. Here, we report that predation risks from either isolated predator or predator odor cues, induce a higher proportion of nymphs to developed into long-winged females among the parent generation, as well as among F1 generation offspring. Surprisingly, these previously threatened long-winged adults survived better when attacked by a predator owing to the enhanced agility level gained from risk experience. The long wing, and increased agility level, provide adaptive benefits for SBPHs to escape from predation and so are able to go on to reproduce.SBPHs responded more strongly to the caged predators (visual + odor risk cues) and predator odor cues, than just the visual cue of the predator. Different risk cues can elicit different levels of responses in prey33,34,35,36. For example, in the case of the Colorado potato beetle, volatile odor cues from the predator stronger reduced the beetle feeding on plants than predator visual and tactile cues35. But a visual cue has been shown to be crucial for insect pollinators detecting and avoiding flowers with predators37. Insect herbivores frequently communicate via chemical odors33,38. Exploiting the odor cue to perceive the presence of predators should have advantages, because the odor cue can be sensed from a long distance and penetrate the blocking effect of foliage or canopy structure39, enabling the prior detection of risks and the preparation of antipredation behaviors.In densely planted rice paddies, the active foraging behavior of rove beetle may serve as a selective pressure favouring the development in SBPHs of a chemical instead of visual pathway to detect the approach of a rove beetle. However, in the F1 generation, the influence of a predator odor cue on the proportion of winged forms was weaker than that of caged predators, indicating the combined effects of odor and visual cues might be stronger than only an odor cue, suggesting that visual cues cannot be ignored. In our experiments, sealed predator cadavers may have weakened the visual cue of the rove beetles, because the lack of motion did not fully represent the normal visual cue.SBPHs frequently exhibit wing plasticity in response to population density and food quality28,29. When nymph density is higher, or food has deteriorated, a higher proportion of macropters will arise28,29. The development of the winged form is thought to be a strategy for SBPHs to emigrate from inhospitable environments. However, we assumed, predation risks could also induce the occurrence of the winged form, because long wings might enable SBPHs to escape from predation. As expected, the results presented here show that a higher proportion of long-winged females and their offspring arose when nymphs or adults were previously exposed to predation risk, demonstrating that SBPHs can express morphologically plastic defenses in response to prior predation risk. Additionally, the higher proportion of wing forms was not only due to the increasing number of winged females (see Fig. 1, the number of winged females in “caged rove beetle” treatment was lower), but also the increasing proportion of winged females among female groups (the decreasing numbers and proportions of wingless females, Fig. 1). To date, similar patterns have only been shown in pea aphids, in which when predation risk (foot prints from lady beetles) is higher during the parent generation, a higher proportion of winged morphs arise in the offspring40,41. In our experiments, we tested the risk effects passing from nymphs to adults and from parents to their offspring with combined risk cues, an odor cue or a visual cue, which better reveals the capacity for flexible defense strategies within SBPHs and the nature of predation risks in the perpetual ‘arms race’ against predation. This is the first example of how insects can express both within- generational and transgenerational morphological plasticity as a defense strategy in response to prior predator threat, and we suggest that this phenomenon is likely to occur more widely.However, SBPHs do not only face a single lethal pressure from their environment as we discussed above. Nymph density, food quality, even the temperatures or photoperiods may play or interplay roles in the induction of wing plasticity in SBPHs28,29. In these situations, the responses of SBPHs may differ from present results, or opposite results can occur. As an example, the growth rates of snails vary depending on snail densities, food supply and the strength of predation risks. Growth rates were higher when snails were reared on high nutrients and in low densities, but decreased steeply as the predation risk increased. Conversely, the growth rate was lower at high densities and with high predation risk, but increased as nutrient availability increased42. As for SBPHs, the proportion of winged adults may be higher if we reared in higher densities combined with high predation risk, or may be lower if the nutrient condition of the rice plants increases (for example, higher fertilizer inputs benefit the development of planthoppers43) and predators are removed. This hypothesis needs to be tested. Further, the rice plant phenotypes (resistant or sensitive phenotypes) are important to the development of planthoppers or leafhoppers44,45,46, and tests of the interactive effects of plant phenotype, plant quality/quantity, nymph density and predation risk on the wing plasticity of SBPHs should provide insights into the evolution of insects within changing environments.Induced transgenerational defense plasticity as shown in SBPHs may be common in many organisms20,47. It allows parents to transfer their risk experience to offspring and promotes their evolutionary fitness20. When SBPH nymphs are exposed to predation risk, they are likely to develop into long-winged females, because it is an advantageous form for them in the current risk environment. However, such predation risk is variable in time and space, and SBPH parents cannot predict when or whether the predators will disappear. Thus, an appropriate strategy to enhance the survival rate of offspring in an unpredictable environment is to continue producing a higher proportion of long-winged forms. Within-generational and transgenerational plasticity of defense should be a successful adaptive defense strategy for SBPHs, given that rove beetle and other groups of predators such as predatory spiders are abundant all around the year in rice paddies.The higher mortality of SBPH nymphs when they experience predation risk, has been broadly addressed before24,48,49. Reduced food intake during risk periods may contribute to this poorer survival outcome, because insects are likely to alter their feeding behavior50,51, or shift from a high-risk host to a safer, but nutritionally inferior, one52, when they detect the presence of predators. However, we did not observe an apparent behavior change in threatened nymphs in our experiments, even those going on to be macropters, compared to the non-threatened ones. For example, changing feeding location, non-feeding related motility, an increase in jump frequency, etc. did not occur in threatened nymphs. Thus, behavior plasticity seems not to be invoked to explain this phenomenon. However, considering the food consumption of sap-sucking SBPHs is difficult to determine, experiments employing electrical penetration graph (EPG) techniques should be conducted to quantify the amounts of sap consumption during risk periods53. This will help to explain whether the higher mortality is due to a change of feeding behavior (less food intake). Furthermore, some obscure internal physiological plasticity may also cause the higher mortality of SBPH nymphs at risk. For example, increased oxidative damage and decreased assimilation efficiency during the risk period may weaken the survival success of SBPH nymphs. Unfortunately, few studies have verified this assumption, although it has been shown that different assimilation efficiencies may arise under predation risk17, or oxidative damage may be induced by predation risk resulting in a slower growth rate54 and decreased escape performance55.SBPHs exhibit sexual differences in both with- and trans- generational morphological plasticity in relation to defense, i.e., threatened nymphs/parents are more likely to develop into long-winged females, due to the different vulnerability of females and males to predation. This predation difference is particularly acute between short-winged females and males, given that the proportion of short-winged females is lower than that seen in control settings (Fig. 1), and we assume the level of vulnerability may depend on their body size and reproductive role. The body sizes of short-winged females are larger than those of long-winged females or males, causing them to be more vulnerable to predation because they are more highly preferred targets for predator. Also, the short-winged female needs to stay and deposit eggs in the bare rice stem, which increases the time window of exposure to predators while, by contrast, long-winged males are slim and are not required to lay eggs, and so should be not be heavily predated. It follows that short-winged females should be more vulnerable to predation than long-winged females or males. Hence, in SHPBs, increasing the proportion of long-wing females in a population creates greater opportunities to migrate to predator-free habitats for reproduction, while at the same time reducing their vulnerability to predation. We hypothesize that the sexual difference in responses should be adaptive, and might be inheritable if predation pressure frequently favors the long-winged forms among populations over multiple generations.Results presented here also show that previously threatened long-winged offspring survived better than previosuly non-threatened ones when attacked by P. fuscipes. Studies suggest prey-altered morphology in response to predation risks should confer a survival advantage (fitness gained), i.e., a better-developed defensive structure13,24, or refuge in having a larger size that increase survival success57. However, wings themselves are without protective functions for SBPHs, as seen in pea aphids41. Thus, we setup behavioral experiments to reveal how threatened long-winged adults may increase their survival when attacked by a predator. Results show threatened long- winged offspring (but not parents) are more active, and respond more quickly, than unthreatened ones, i.e., a higher number of attacks are needed for P. fuscipes to capture a previously threatened long-winged offspring than one that has not been threatened before. We suggest the increased agility level is not because of the long wing itself, but due to the enhanced muscle strength in the legs of long-winged adults, because in our observation, long-winged adults avoid attack mainly by jumping but not by flight, probably because a jump needs less reaction time than flight.We only observed transgenerational plasticity of induced behavioral defense in SBPHs. This generational difference (within- and trans-generational) in behavioral defense in SBPHs may reflect potential carry-over effects from parents. To our knowledge, the generational difference in defense has rarely been shown in insects, though in pea aphids a fluctuating expression of transgenerational defensive traits (long wing) over generations when predation risk was present or absent has been reported58. We also expect there will be cumulative effects59 accumulated by SBPHs from the parent generation to the F1 generation. However, we are not certain whether these effects exist in our experiments. To determine this, experiments examining defensive traits across multigeneration should be conducted.However, if predation risk increases the number of agile, long-winged SBPH adults, which are of benefit in respect of dispersal, migration, and thus spreading rice viruses, the application of P. fuscipes in biological control appears ultimately to weaken the control effectiveness. Also, a study with field experiments found that predatory ladybugs increase the number of dispersed aphid nymphs, especially in plants with lower resistance. However, surprising results show that the higher number of dispersed aphid nymphs will not necessarily translate into population growth because dispersed aphids are weak (less food intake) and more easily predated by predators60. Thus, the benefits of anti-predator defense in aphids will, over time, translate into negative developmental costs that suppress the aphid population. As for SBPHs, threatened long-winged females perform well in dispersal and defense, but worse in development and reproduction. Recent experiments reveal that previously threatened long-winged females have a longevity that is three days shorter, and produces about 60 fewer eggs per female, than non-threatened long-winged females (unpublished data). Consequently, these negative effects would eventually translate into lower population growth rates within SBPHs. Thus, the introduction of the predation risk from P. fuscipes to control SBPHs is workable, since field experiments in controlling western flower thrips and grasshoppers by exposure to predation risk have been successful49,61, and the main purpose of biological control is to suppress the pest population beneath the relevant economic threshold, and reduce plant mass loss without necessarily eliminating the pest altogether.This study advances the importance of predation risk on the induction of flexible anti-predation defenses in insect parents and their offspring, uncovers the mediating mechanisms, shows how this anti-predation defense expresses differently between sexes, and further explores the adaptation significance of these defense traits for insects exposed to unpredictable environments. These findings should prove important for predicting SBPH migration or dispersal, conducting effective pest control, and better understanding prey-predator interactions. However, future work should examine the effects of predation risks from other groups of predators or parasites on the physiological and behavioral plasticity of SBPHs. More

  • in

    Great Barrier Reef: accept ‘in danger’ status, there’s more to gain than lose

    WORLD VIEW
    18 August 2021

    Great Barrier Reef: accept ‘in danger’ status, there’s more to gain than lose

    The Australian government must embrace UNESCO’s assessment to marshal the resources needed to protect the unique coral ecosystem.

    Tiffany H. Morrison

    0

    Tiffany H. Morrison

    Tiffany H. Morrison is a political geographer specializing in marine interventions at James Cook University in Townsville, Australia.

    View author publications

    You can also search for this author in PubMed
     Google Scholar

    Share on Twitter
    Share on Twitter

    Share on Facebook
    Share on Facebook

    Share via E-Mail
    Share via E-Mail

    Download PDF

    No one denies the cascade of climate-induced coral bleaching that devastated huge portions of the Great Barrier Reef in 2016, nor the subsequent bleaching. No one questions the Queensland government’s 2019 report (see go.nature.com/3ckg) that the reef’s condition near the shore is poor.Yet last month, the World Heritage Committee of the United Nations organization UNESCO caved to lobbying from the Australian government — pressured by fossil-fuel, agricultural and mining interests — and kept the Great Barrier Reef off its list of ecosystems ‘in danger’. In my view, this decision is wrong, factually and strategically. It leaves both UNESCO and Australia weaker against the climate crisis.I study the governance of approximately 250 ecosystems with World Heritage status because of their outstanding value to humanity — including attempts to curtail runaway industrial development of Vietnam’s Ha Long Bay and overzealous urbanization along Florida’s Everglades wetlands.There are benefits to an in-danger listing: the Belize Barrier Reef Reserve System was placed on the list in 2009. The World Heritage Fund then provided technical and financial assistance for its restoration. By 2018, mangrove coverage was restored nearly to 1996 levels, with clearing in protected areas almost entirely curtailed. The whole maritime zone was under a moratorium on oil and gas production. Restoration work is ongoing, but the Belize reef is no longer on the list.
    Save reefs to rescue all ecosystems
    This July, UNESCO proposed to list the Great Barrier Reef as in danger owing to severe coral bleaching, poor water quality and inaction on climate change.In arguing against the listing, the Australian government did not directly deny the reef’s parlous state, but did play down its condition. The government also argued that the listing would decrease tourism revenues, that Australia had too little time to respond and should not be held responsible for global change, and that UNESCO should not supersede national sovereignty on climate-change policy.Australian environment minister Sussan Ley lobbied committee members from more than a dozen countries to override UNESCO’s recommendation. Australia avoided an in-danger listing in 2015 using similar tactics and by touting a sustainability plan. The following year saw the worst coral bleaching in the world’s history.But changes are in the wind. After back-to-back coral bleaching in 2016–17 and the tragic 2020 bush fires, more Australian voters, industries and even conservative politicians are calling for strong efforts against climate change.Accepting an in-danger listing for the reef could tip the balance past gridlock. More than 70% of Australians think that formally acknowledging the reef’s endangered state would spur action. In 1993, former US president Bill Clinton’s administration requested that UNESCO certify Florida’s Everglades as in danger. This helped to bring industry opponents on board to better manage coastal development. Had the Great Barrier Reef been listed as in danger in 2015, fossil-fuel developments in the catchment areas draining into the reef would have struggled to get approval.Australia’s most conservative politicians will argue that avoiding an in-danger listing in 2022 is necessary to boost economic development. But this will embarrass Australia later. As more marine heating occurs globally, Australia will struggle to defend its inaction on climate to the UN climate-change conference in November and to the World Heritage Committee next year. Even the Queensland Tourism Industry Council has said keeping the reef’s status under the spotlight is a “call to the world to do more on climate change”.
    Fevers are plaguing the oceans — and climate change is making them worse
    And undercutting the listing undermines the purpose of the World Heritage Committee. Since 1972, 41 ecosystems have been considered for the in-danger list — 27 of them more than once — but not officially inscribed, even though UNESCO and its advisory body had assessed these ecosystems as threatened, or more threatened than those already listed. The number of sites on the list has declined by almost one-third since 2001, although threats continue to grow and there are more ecosystems on the overall World Heritage List.However, destabilizing strategies are mainly due to a small group of nations — including countries in the Organisation for Economic Co-operation and Development, such as Australia and Spain. World Heritage status and in-danger listings often work as intended: the managers of 73% of sites do comply with their responsibilities.Concerned observers are helping the World Heritage Committee to protect itself from political manipulation. In February 2020, a consortium of 76 organizations and individuals petitioned UNESCO to consider climate change in its World Heritage decisions. A nascent international network known as World Heritage Watch hopes to provide more oversight and monitoring of self-interested states. Ecologists and non-profit organizations are using remote sensing and citizen science to track and expose degradation of protected areas (see go.nature.com/2xn1) and hold governments accountable.UNESCO and its World Heritage Committee grasp the stakes. A new draft policy clearly states that climate-related degradation of a World Heritage Area can be used as the basis for in-danger listing; it will probably be ratified later this year at the UNESCO General Assembly. This policy will shine a harsh light on the intensifying geopolitics of climate change. Advanced economies, such as Australia, with high per-capita emissions but limited climate action, will need to find alternative ways to protect resources and jobs.

    Nature 596, 319 (2021)
    doi: https://doi.org/10.1038/d41586-021-02220-3

    Related Articles

    Fevers are plaguing the oceans — and climate change is making them worse

    Save reefs to rescue all ecosystems

    Subjects

    Climate change

    Conservation biology

    Government

    Latest on:

    Climate change

    ‘Polluter pays’ policy could speed up emission reductions and removal of atmospheric CO2
    News & Views 16 AUG 21

    Warming world, women in science — the week in infographics
    News 13 AUG 21

    IPCC climate report: Earth is warmer than it’s been in 125,000 years
    News 09 AUG 21

    Government

    Brazilian road proposal threatens famed biodiversity hotspot
    News 17 AUG 21

    The world must cooperate to avoid a catastrophic space collision
    Editorial 11 AUG 21

    From the archive
    News & Views 10 AUG 21

    Jobs

    Locum Associate or Senior Editor, Immunology

    Springer Nature
    London, United Kingdom

    Position title: Postdoctoral Research Fellows – Air Force Science & Technology Fellowship Program

    National Academies of Sciences, Engineering, and Medicine
    Various locations in the U.S., United States

    Postdoctoral Research Associates – NRC Research Associateship Programs

    National Academies of Sciences, Engineering, and Medicine
    Various locations in the U.S., United States

    Research Associate in the Centre for Developmental Neurobiology

    King’s College London (KCL)
    London, United Kingdom

    Nature Briefing
    An essential round-up of science news, opinion and analysis, delivered to your inbox every weekday.

    Email address

    Yes! Sign me up to receive the daily Nature Briefing email. I agree my information will be processed in accordance with the Nature and Springer Nature Limited Privacy Policy.

    Sign up More

  • in

    Intermediate ice scour disturbance is key to maintaining a peak in biodiversity within the shallows of the Western Antarctic Peninsula

    1.Dell, J. et al. Interaction diversity maintains resiliency in a frequently disturbed ecosystem. Front. Ecol. Evol. 7, 145 (2019).Article 

    Google Scholar 
    2.White, P. S. & Pickett, S. T. A. In The Ecology of Natural Disturbance and Patch Dynamics (eds S. T. A. Pickett & P. S. White) 3–13 (Academic Press, 1985).3.Newman, E. A. Disturbance ecology in the anthropocene. Front. Ecol. Evolut. https://doi.org/10.3389/fevo.2019.00147 (2019).Article 

    Google Scholar 
    4.Barnosky, A. D. et al. Approaching a state shift in Earth’s biosphere. Nature 486, 52–58 (2012).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    5.Yuan, Z., Jiao, F., Li, Y. & Kallenbach, R. L. Anthropogenic disturbances are key to maintaining the biodiversity of grasslands. Sci. Rep. 6, 22132 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    6.Hughes, A. R., Byrnes, J. E., Kimbro, D. L. & Stachowicz, J. J. Reciprocal relationships and potential feedbacks between biodiversity and disturbance. Ecol. Lett. 10, 849–864. https://doi.org/10.1111/j.1461-0248.2007.01075.x (2007).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    7.Connell, J. H. & Slatyer, R. O. Mechanisms of succession in natural communities and their role in community stability and organization. Am. Nat. 111, 1119–1144 (1977).Article 

    Google Scholar 
    8.Connell, J. H. Diversity in tropical rain forests and coral reefs. Science 199, 1302–1310 (1978).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    9.Fox, J. W. The intermediate disturbance hypothesis should be abandoned. Trends Ecol. Evol. 28, 86–92. https://doi.org/10.1016/j.tree.2012.08.014 (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    10.Sheil, D. & Burslem, D. F. Disturbing hypotheses in tropical forests. Trends Ecol. Evol. 18, 18–26 (2003).Article 

    Google Scholar 
    11.Teixidó, N., Garrabou, J., Gutt, J. & Arntz, W. E. Recovery in Antarctic benthos after iceberg disturbance: Trends in benthic composition, abundance and growth forms. Mar. Ecol. Prog. Ser. 278, 1–16. https://doi.org/10.3354/meps278001 (2004).ADS 
    Article 

    Google Scholar 
    12.Teixidó, N., Garrabou, J., Gutt, J. & Arntz, W. Iceberg disturbance and successional spatial patterns: the case of the shelf Antarctic benthic communities. Ecosystems 10, 143–158 (2007).Article 

    Google Scholar 
    13.Johst, K., Gutt, J., Wissel, C. & Grimm, V. Diversity and disturbances in the Antarctic megabenthos: Feasible versus theoretical disturbance ranges. Ecosystems 9, 1145–1155 (2006).Article 

    Google Scholar 
    14.Mackey, R. L. & Currie, D. J. The diversity-disturbance relationship: Is it generally strong and peaked?. Ecology 82, 3479–3492. https://doi.org/10.1890/0012-9658(2001) (2001).Article 

    Google Scholar 
    15.Huston, M. A. Disturbance, productivity, and species diversity: Empiricism vs. logic in ecological theory. Ecology 95, 2382–2396. https://doi.org/10.1890/13-1397.1 (2014).Article 

    Google Scholar 
    16.Smale, D. A., Brown, K. M., Barnes, D. K., Fraser, K. P. & Clarke, A. Ice scour disturbance in Antarctic waters. Science 321, 371. https://doi.org/10.1126/science.1158647 (2008).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    17.Griffiths, H. J., Danis, B. & Clarke, A. Quantifying Antarctic marine biodiversity: The SCAR-MarBIN data portal. Deep Sea Res. Part II 58, 18–29. https://doi.org/10.1016/j.dsr2.2010.10.008 (2011).ADS 
    Article 

    Google Scholar 
    18.Grange, L. J. & Smith, C. R. Megafaunal communities in rapidly warming fjords along the West Antarctic Peninsula: Hotspots of abundance and beta diversity. PLoS ONE 8, e77917 (2013).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    19.Gutt, J., Griffiths, H. J. & Jones, C. D. Circumpolar overview and spatial heterogeneity of Antarctic macrobenthic communities. Mar. Biodivers. 43, 481–487. https://doi.org/10.1007/s12526-013-0152-9 (2013).Article 

    Google Scholar 
    20.Potthoff, M., Johst, K. & Gutt, J. How to survive as a pioneer species in the Antarctic benthos: Minimum dispersal distance as a function of lifetime and disturbance. Polar Biol. 29, 543–551 (2006).Article 

    Google Scholar 
    21.Convey, P. et al. The spatial structure of Antarctic biodiversity. Ecol. Monogr. 84, 203–244 (2014).Article 

    Google Scholar 
    22.Peck, L. S., Brockington, S., Vanhove, S. & Beghyn, M. Community recovery following catastrophic iceberg impacts in a soft-sediment shallow-water site at Signy Island, Antarctica. Mar. Ecol Progr. Ser. 186, 1–8 (1999).ADS 
    Article 

    Google Scholar 
    23.Lee, H., Vanhove, S., Peck, L. & Vincx, M. Recolonisation of meiofauna after catastrophic iceberg scouring in shallow Antarctic sediments. Polar Biol. 24, 918–925. https://doi.org/10.1007/s003000100300 (2001).Article 

    Google Scholar 
    24.Armstrong, T. World Meteorological Organization. WMO sea-ice nomenclature. Terminology, codes and illustrated glossary. Edition 1970. Geneva, Secretariat of the World Meteorological Organization, 1970. [ix], 147 p. [including 175 photos]+ corrigenda slip. (WMO/OMM/BMO, No. 259, TP. 145.). J. Glaciol. 11, 148–149 (1972).25.Robinson, B. J., Barnes, D. K. & Morley, S. A. Disturbance, dispersal and marine assemblage structure: A case study from the nearshore Southern Ocean. Mar. Environ. Res. 160, 105025 (2020).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    26.Gutt, J., Starmans, A. & Dieckmann, G. Impact of iceberg scouring on polar benthic habitats. Mar. Ecol. Prog. Ser. 137, 311–316 (1996).ADS 
    Article 

    Google Scholar 
    27.Barnes, D. K. A. & Conlan, K. E. Disturbance, colonization and development of Antarctic benthic communities. Philos. Trans. R. Soc. Lond. B Biol. Sci. 362, 11–38. https://doi.org/10.1098/rstb.2006.1951 (2007).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    28.Smale, D. A. Ecological traits of benthic assemblages in shallow Antarctic waters: Does ice scour disturbance select for small, mobile, secondary consumers with high dispersal potential?. Polar Biol. 31, 1225–1231. https://doi.org/10.1007/s00300-008-0461-9 (2008).Article 

    Google Scholar 
    29.Barnes, D. K. A. The influence of ice on polar nearshore benthos. J. Mar. Biol. Assoc. U.K. 79, 401–407 (1999).Article 

    Google Scholar 
    30.Gutt, J. On the direct impact of ice on marine benthic communities, a review. Polar Biol. 24, 553–564 (2001).Article 

    Google Scholar 
    31.Barnes, D. K. A. & Tarling, G. A. Polar oceans in a changing climate. Curr. Biol. 27, R454–R460. https://doi.org/10.1016/j.cub.2017.01.045 (2017).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    32.Barnes, D. K. A., Fleming, A., Sands, C. J., Quartino, M. L. & Deregibus, D. Icebergs, sea ice, blue carbon and Antarctic climate feedbacks. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 376, 20170176. https://doi.org/10.1098/rsta.2017.0176 (2018).ADS 
    Article 

    Google Scholar 
    33.Cook, A. J., Fox, A. J., Vaughan, D. G. & Ferrigno, J. G. Retreating glacier fronts on the Antarctic Peninsula over the past half-century. Science 308, 541–544. https://doi.org/10.1126/science.1104235 (2015).ADS 
    CAS 
    Article 

    Google Scholar 
    34.Cook, A. et al. Ocean forcing of glacier retreat in the western Antarctic Peninsula. Science 353, 283–286 (2016).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    35.Clarke, A. et al. Climate change and the marine ecosystem of the western Antarctic Peninsula. Philos. Trans. R. Soc. Lond. B Biol. Sci. 362, 149–166. https://doi.org/10.1098/rstb.2006.1958 (2007).Article 
    PubMed 

    Google Scholar 
    36.Turner, J. & Comiso, J. Solve Antarctica’s sea-ice puzzle. Nat. News 547, 275 (2017).CAS 
    Article 

    Google Scholar 
    37.Meredith, M. P. & King, J. C. Rapid climate change in the ocean west of the Antarctic Peninsula during the second half of the 20th century. Geophys. Res. Lett. https://doi.org/10.1029/2005GL024042 (2005).Article 

    Google Scholar 
    38.Barnes, D. K. A. & Souster, T. Reduced survival of Antarctic benthos linked to climate-induced iceberg scouring. Nat. Clim. Chang. 1, 365–368. https://doi.org/10.1038/nclimate1232 (2011).ADS 
    Article 

    Google Scholar 
    39.Parkinson, C. L. Global sea ice coverage from satellite data: Annual cycle and 35-yr trends. J. Clim. 27, 9377–9382. https://doi.org/10.1175/jcli-d-14-00605.1 (2014).ADS 
    Article 

    Google Scholar 
    40.Rogers, A. et al. Antarctic futures: An assessment of climate-driven changes in ecosystem structure, function, and service provisioning in the Southern Ocean. Ann. Rev. Mar. Sci. 12, 87–120 (2020).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    41.Morley, S. A. et al. Global drivers on Southern Ocean ecosystems: Changing physical environments and anthropogenic pressures in an Earth system. Front. Mar. Sci. 7, 1097 (2020).Article 

    Google Scholar 
    42.Barnes, D. K. et al. Blue carbon gains from glacial retreat along Antarctic fjords: What should we expect?. Glob. Change Biol. 26, 2750–2755 (2020).ADS 
    Article 

    Google Scholar 
    43.Barnes D. K. A. Blue carbon on polar and subpolar seabeds. In Carbon capture, utilization and sequestration (InTech, 2018). https://doi.org/10.5772/intechopen.78237.44.Bowler, D. et al. The geography of the Anthropocene differs between the land and the sea. bioRxiv https://doi.org/10.1101/432880 (2019).Article 

    Google Scholar 
    45.Arntz, W., Brey, T. & Gallardo, V. Antarctic zoobenthos. Oceanogr. Mar. Biol. 32, 241–304 (1994).
    Google Scholar 
    46.Clarke, A. Marine benthic populations in Antarctica: Patterns and processes. Antarct. Res. Ser. 70, 373–388 (1996).Article 

    Google Scholar 
    47.Fillinger, L., Janussen, D., Lundälv, T. & Richter, C. Rapid glass sponge expansion after climate-induced Antarctic ice shelf collapse. Curr. Biol. 23, 1330–1334 (2013).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    48.Clarke, A., Meredith, M. P., Wallace, M. I., Brandon, M. A. & Thomas, D. N. Seasonal and interannual variability in temperature, chlorophyll and macronutrients in northern Marguerite Bay, Antarctica. Deep Sea Res. Part II 55, 1988–2006. https://doi.org/10.1016/j.dsr2.2008.04.035 (2008).ADS 
    Article 

    Google Scholar 
    49.Barnes, D. K. A. Iceberg killing fields limit huge potential for benthic blue carbon in Antarctic shallows. Glob. Chang. Biol. 23, 2649–2659. https://doi.org/10.1111/gcb.13523 (2017).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    50.Pinkerton, M., Bradford-Grieve, J., Bowden, D. & Cummings, V. Benthos: Trophic modelling of the Ross Sea. Support. Docum. CCAMLR Sci. 17, 1–31 (2010).
    Google Scholar 
    51.Pielou, E. Shannon’s formula as a measurement of species diversity: It’s use and disuse. Am. Nat. 100, 463–465 (1966).Article 

    Google Scholar 
    52.Fisher, R. A., Corbet, A. S. & Williams, C. B. The relation between the number of species and the number of individuals in a random sample of an animal population. J. Anim. Ecol. 1, 42–58 (1943).Article 

    Google Scholar 
    53.Everitt, B. & Skrondal, A. The Cambridge Dictionary of Statistics Vol. 106 (Cambridge University Press, Cambridge, 2002).MATH 

    Google Scholar 
    54.Smale, D. A., Barnes, D. K. A. & Fraser, K. P. P. The influence of ice scour on benthic communities at three contrasting sites at Adelaide Island, Antarctica. Aust. Ecol. 32, 878–888. https://doi.org/10.1111/j.1442-9993.2007.01776.x (2007).Article 

    Google Scholar 
    55.Peck, L. S., Convey, P. & Barnes, D. K. A. Environmental constraints on life histories in Antarctic ecosystems: Tempos, timings and predictability. Biol. Rev. 81, 75–109. https://doi.org/10.1017/s1464793105006871 (2006).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    56.Waller, C., Worland, M., Convey, P. & Barnes, D. Ecophysiological strategies of Antarctic intertidal invertebrates faced with freezing stress. Polar Biol. 29, 1077–1083 (2006).Article 

    Google Scholar 
    57.Barnes, D. K. A. Polar zoobenthos blue carbon storage increases with sea ice losses, because across-shelf growth gains from longer algal blooms outweigh ice scour mortality in the shallows. Glob. Chang Biol. 23, 5083–5091. https://doi.org/10.1111/gcb.13772 (2017).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    58.Smith, C. R., Mincks, S. & DeMaster, D. J. A synthesis of bentho-pelagic coupling on the Antarctic shelf: Food banks, ecosystem inertia and global climate change. Deep Sea Res. Part II 53, 875–894 (2006).ADS 
    Article 

    Google Scholar 
    59.Jansen, J. et al. Abundance and richness of key Antarctic seafloor fauna correlates with modelled food availability. Nat. Ecol. Evolut. 2, 71–80 (2018).Article 

    Google Scholar 
    60.Henley, S. F. et al. Changing biogeochemistry of the Southern Ocean and its ecosystem implications. Front. Mar. Sci. 7, 581 (2020).Article 

    Google Scholar 
    61.Marshall, G. J. et al. Causes of exceptional atmospheric circulation changes in the Southern Hemisphere. Geophys. Res. Lett. 31, 14 (2004).Article 

    Google Scholar 
    62.Ashton, G. V., Morley, S. A., Barnes, D. K., Clark, M. S. & Peck, L. S. Warming by 1 C drives species and assemblage level responses in Antarctica’s marine shallows. Curr. Biol. 27, 2698-2705e2693 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    63.Riesgo, A. et al. Some like it fat: Comparative ultrastructure of the embryo in two demosponges of the genus Mycale (order poecilosclerida) from Antarctica and the Caribbean. PLoS ONE 10, e0118805 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    64.Toszogyova, A. & Storch, D. Global diversity patterns are modulated by temporal fluctuations in primary productivity. Glob. Ecol. Biogeogr. 28, 1827–1838 (2019).Article 

    Google Scholar 
    65.Clark, G. F. et al. Light-driven tipping points in polar ecosystems. Glob. Change Biol. 19, 3749–3761 (2013).ADS 
    Article 

    Google Scholar 
    66.Brockington, S., Clarke, A. & Chapman, A. Seasonality of feeding and nutritional status during the austral winter in the Antarctic sea urchin Sterechinus neumayeri. Mar. Biol. 139, 127–138 (2001).Article 

    Google Scholar 
    67.Fratt, D. B. & Dearborn, J. Feeding biology of the Antarctic brittle star Ophionotus victoriae (Echinodermata: Ophiuroidea). Polar Biol. 3, 127–139 (1984).Article 

    Google Scholar 
    68.Sahade, R., Tatián, M. & Esnal, G. B. Reproductive ecology of the ascidian Cnemidocarpa verrucosa at Potter Cove, South Shetland Islands, Antarctica. Mar. Ecol. Progr. Ser. 272, 131–140 (2004).ADS 
    Article 

    Google Scholar 
    69.Dayton, P. K. et al. Recruitment, growth and mortality of an Antarctic hexactinellid sponge, Anoxycalyx joubini. PLoS ONE 8, e56939 (2013).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    70.Vacchi, M., Cattaneo-Vietti, R., Chiantore, M. & Dalù, M. Predator-prey relationship between the nototheniid fish Trematomus bernacchii and the Antarctic scallop Adamussium colbecki at Terra Nova Bay (Ross Sea). Antarct. Sci. 12, 64–68 (2000).ADS 
    Article 

    Google Scholar 
    71.Sheil, D. & Burslem, D. F. Defining and defending Connell’s intermediate disturbance hypothesis: a response to Fox. Trends Ecol. Evol. 28, 571–572. https://doi.org/10.1016/j.tree.2013.07.006 (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar  More

  • in

    Environmental stress leads to genome streamlining in a widely distributed species of soil bacteria

    A. Strain sampling and isolationBradyrhizobium is a commonly occurring genus in soil [21]. Closely related Bradyrhizobium diazoefficiens (previously Bradyrhizobium japonicum) strains were isolated from soil, as previously described [20, 22]. In brief, Bradyrhizobium isolates that formed symbiotic associations with a foundational legume species in the censused region, Acacia acuminata, were isolated from soil sampled along a large region spanning ~300,000 km2 in South West Australia, a globally significant biodiversity hotspot [23]. In total 60 soil samples were collected from twenty sites (3 soil samples per site; Supplementary Fig. S1) and 380 isolates were sequenced (19 isolates per site, 5 or 6 isolates per soil sample, each isolate re-plated from a single colony at least 2 times). Host A. acuminata legume plants were inoculated with field soil in controlled chamber conditions and isolates were cultured on Mannitol Yeast agar plates from root nodules (see [20, 22] for details). A total of 374 strains were included in this study after removing 5 contaminated samples and one sample that was a different Bradyrhizobium species; non- Bradyrhizobium diazoefficiens sample removal was determined from 16S rRNA sequences extracted from draft genome assemblies (Method C) using RNAmmer [24].B. Environmental variation among sampled sitesIn this study, I focus on environmental factors (temperature, rainfall, soil pH and salinity) previously identified to impact either rhizobia growth performance, functional fitness or persistence in soil [25,26,27,28] and where a directionality of rhizobial stress response could be attributed with respect to environmental variation present in the sampled region (i.e. stress occurs at high temperatures, low rainfall, high acidity and high salinity). Each environmental factor was standardised to a mean of 0 and a standard deviation of 1, and pH and rainfall scales were reversed to standardise stress responses directions so that low stress is at low values and high stress is at high values for all factors (Supplementary Fig. S2). Additionally, salinity was transformed using a log transformation (log(x + 0.01) to account for some zeroes) prior to standardisation.C. Isolate sequencing and pangenome annotationIllumina short reads (150 bp paired-end) were obtained and draft genome assemblies were generated using Unicycler from a previous study [29]. Resulting assemblies were of good assembly quality (99.2% of all strains had >95.0% genome completeness score according to BUSCO [30]; Table S1; assembled using reads that contained nominal 0.016 ± 0.00524% non-prokaryotic DNA content across all 374 isolates, according to Kraken classification [31]). Protein coding regions (CDS regions) were identified using Prokka [32] and assembled into a draft pangenome using ROARY [33], which produced a matrix of counts of orthologous gene clusters (i.e. here cluster refers to a set of protein-coding sequences containing all orthologous variants from all the different strains, grouped together and designated as a single putative gene). Gene clusters that occurred in 99% of strains were designated as ‘core genes’ and used to calculate the ‘efficiency of selection’ [34, 35] (measured as dN/dS, Method G.2) and population divergence measured as Fixation Index ‘Fst’, Method H) across each environmental stress factor. The identified gene clusters were then annotated using eggNOG-mapper V2 [36] and the strain by gene cluster matrix was reaggregated using the Seed ortholog ID returned by eggNOG-mapper as the protein identity. Out of the total 2,744,533 CDS regions identified in the full sample of 374 strains, eggNOG-mapper was able to assign 2,612,345 of them to 91,230 unique Seed orthologs. These 91,230 protein coding genes constituted the final protein dataset for subsequent analyses.D. Calculation and statistical analysis of gene richness and pangenome diversity along the stress gradientGene richness was calculated as the total number of unique seed orthologues for each strain (i.e. genome). Any singleton genes that occurred in only a single strain, as well as ‘core’ genes that occurred in every strain (for symmetry, and because these are equally uninformative with respect to variation between strains) were removed, leaving 74,089 genes in this analysis. Gene richness (being count data) was modelled on a negative binomial distribution (glmer.nb function) as a function of each of the four environmental stressors as predictors using the lme4 package in R [37], also accounting for hierarchical structure in the data by including site and soil sample as random effects.To rule out potentially spurious effects of assembly quality (i.e. missed gene annotations due to incomplete draft genomes) on key findings, I confirmed no significant association between gene richness and genome completeness (r = 0.042, p = 0.4224, Fig. S3).Finally, pangenome diversity was calculated as the total number of unique genes that occurred in each soil sample (since multiple strains were isolated from a single soil sample). Pangenome diversity was modelled the same as gene richness, except here soil sample was not included as a random effect.E. Calculation of network and duplication traits for each geneI used the seed orthologue identifier from eggNOG-mapper annotations to query matching genes within StringDB ([38]; https://string-db.org/), which collects information on protein-protein interactions. Out of 91,230 query seed orthologues, 73,126 (~80%) returned a match in STRING. For matching seed orthologue hits, a network was created by connecting any proteins that were annotated as having pairwise interactions in the STRING database using the igraph package in R [39]. Two vertex-based network metrics were calculated for each gene: betweenness centrality, which measures a genes tendency to connect other genes in the gene network, and mean cosine similarity, which is a measure of how much a gene’s links to other genes are similar to other genes.Betweenness centrality was calculated using igraph (functional betweenness). For mean cosine similarity, a pairwise cosine similarity was first calculated between all genes. To do this, the igraph network object was converted into a (naturally sparse yet large) adjacency matrix and the cosSparse function in qlcMatrix in R [40] was used to calculate cosine similarity between all pairs of genes. To obtain an overall cosine similarity trait value for each gene, the average pairwise cosine similarity to all other genes in the network was calculated.Finally, gene duplication level was calculated for each gene as one additional measure of ‘redundancy’, by calculating the average number of gene duplicates found within the same strain. Duplicates were identified as CDS regions with the same Seed orthologue ID.F. Gene distribution modelsTo determine how gene traits predict accessory genome distributions patterns along the stress gradients, I first calculated a model-based metric (hereafter and more specifically a standardised coefficient, ‘z-score’) of the relative tendency of each gene to be found in different soil samples across the four stress gradients (heat, salinity, acidity, and aridity). This was achieved by modelling each gene’s presence or absence in a strain as a function of the four stress gradients, with site and soil sample as a random effect, using a binomial model in lme4 (the structure of the model being the same as the gene richness model, only the response is different). To reduce computational overhead, these models were only run for the set of genes that were used in the gene richness analysis (e.g. after removing singletons and core genes), and which had matching network traits calculated (e.g. they occurred in the STRING database; n = 64,867 genes). Distribution models were run in tandem for each gene using the manyany function in the R package mvabund [41]. Standardised coefficients, or z-scores (coefficient/standard error) indicating the change in the probability of occurrence for each gene across each of the stress gradients were extracted. More negative coefficients correspond to genes that are more likely to be absent in high stress (and vice versa for positive coefficients).To determine how network and duplication traits influence the distribution of genes across the stress gradient, I performed a subsequent linear regression model where the gene’s z-scores was the response and gene traits as predictors. The environmental stress type (i.e. acidity, aridity, heat and salinity) was included as a categorical predictor, and the interaction between stress category and gene function traits were used to infer the influence of gene function traits on gene distributions in a given stress type (see Supplementary Methods S1 for z-score transformation).G. Quantifying molecular signals of natural selection on accessory and core genesTo examine molecular signatures of selection in accessory and core genes, I calculated dN/dS for a subsample of the total pool (n=74,089 genes), which estimates the efficiency of selection [34, 35]. Two major questions relevant to dN/dS that are addressed here require a different gene subsampling approach:(1) Do variable environmental stress responses lead to different dN/dS patterns among accessory genes?Here, I subsampled accessory genes (total accessory gene pool across 374 strains, 74,089) to generate and compare dN/dS among 3 categorical groups, each representing a different level of stress response based on their z-scores (accessory genes that either strongly increase, decrease or have no change in occurrence as stress increases; n = 1000 genes/category; see Supplementary Methods S2 for subsample stratification details).For each gene, sequences were aligned using codon-aware alignment tool, MACSE v2 [42]. dN/dS was estimated by codon within each gene using Genomegamap’s Bayesian model-based approach [43], which is a phylogeny-free method optimised for within bacterial species dN/dS calculation (see Supplementary Methods S3 for dN/dS calculation/summarisation; S9 for xml configuration). The proportion of codons with dN/dS that were credibly less than 1 (purifying selection) and those credibly greater than 1 (positive selection) were analysed, with respect to the genes’ occurrence response to stress. Specifically, I modelled the proportion of codons with dN/dS  1 was overall too low to analyse in this way, so the binary outcome (a gene with any codons with dN/dS  > 1 or not) was modelled using a binomial response model with the response categories as predictors (see Supplementary Methods S4 for details of both models).(2) Does dN/dS among microbial populations vary across environmental stress?Here, I compared the average change in dN/dS in core genes present across all environments at the population level (i.e. all isolates from one soil sample), which is used here to measure the change in the efficiency of selection across each stress gradient. Core genes were used since they occur in all soil samples, allowing a consistent set and sample size of genes to be used in the population-level dN/dS calculation. This contrasts with the previous section, which quantifies gene-level dN/dS on extant accessory genes that intrinsically have variable presence or absence across soil samples. For computational feasibility, 500 core genes were subsampled (total core 1015 genes) and, for each gene, individual strain variants were collated into a single fasta file based on soil sample membership, such that dN/dS could be calculated separately for each gene within each soil sample (i.e. 60 soil samples × 500 genes = 30,000 fasta files). Each fasta file was then aligned in MACSE and dN/dS calculated using the same methodology for accessory genes (Supplementary Method S3). This enabled the average dN/dS in a sample to be associated with soil-sample level environmental stress variables. Specifically, I modelled the mean proportion of codons with dN/dS  1 due to overall rarity of positive selection (average proportion of genes where at least 1 codon with dN/dS  > 1 was ~0.006). This low level of positive selection is expected for core genes which tend to be under strong selective constraint.H. Calculation and analysis of Fixation index (Fst) along stress gradientsUsing the core genome alignment (all SNPs among 1015 core genes) generated previously with ROARY, I computed pairwise environmentally-stratified Fst. Each soil sample (n = 60) was first placed into one of 5 bins based on their distances in total environmental stress space (using all four stress gradients), with the overall goal of generating roughly evenly sized bins such that the environmental similarity of stress was greater within bins than between (see Supplementary Methods S6 and Fig. S4 for clustering algorithm details). Next, fasta alignments were converted to binary SNPs using the adegenet package. Pairwise Fst between all strains (originating from a particular soil sample) within a single bin was calculated using StAMPP in R [44]. For each strain pair, the average of the two stress gradient values was assigned.The relationship between pairwise Fst and the average stress value was evaluated using a linear regression model with each of the four stress values as predictors. Since the analysis uses pairwise data (thus violating standard independence assumptions), the significance of the relationship was determined using a permutation test (see Supplementary Methods S7 for details).I. Chromosomal structure analysis of gene loss patternsTo gain insight into structural variation and test for regional hotspots in gene loss along the chromosome, I mapped each gene’s stress response (i.e. probability of loss or gain indicated by each genes z-score) onto a completed Bradyrhizobium genome (strain ‘36_1’ from the same set of 374 strains (Genbank CP067102.1; [45]). Putative CDS positions on the complete genome were determined using Prokka and annotated with SEED orthologue ID’s using eggNOG-mapper. Matching accessory genes derived from the full set of 374 incomplete draft genomes (n = 74,089) were mapped to positions on the complete genome (6274 matches). The magnitude of gene loss or gain (as measured by their standardised z-scores for each environment from the gene distribution models; see Method F) was then modelled across the genome using a one-dimensional spatial smoothing model. This model was implemented in R INLA [46] (www.r-inla.org), and models a response in a one-dimensional space using a Matern covariance-based random effect. The method uses an integrated nested Laplace approximation to a Bayesian posterior distribution, with a cyclical coordinate system to accommodate circular genomes. The model accounts for spatial non-independence among sites and produces a continuous posterior distribution of average z-score predictions along the entire genome, which was then used to visualise potential ‘hotspots’ of gene loss or gain. The modelling procedure was repeated, instead with gene network traits, such that model predictions of similarity and betweenness could be visualised on the reference chromosome. More