More stories

  • in

    Small-scale population divergence is driven by local larval environment in a temperate amphibian

    Aguillon SM, Fitzpatrick JW, Bowman R, Schoech SJ, Clark AG, Coop G et al. (2017) Deconstructing isolation-by-distance: the genomic consequences of limited dispersal. PLOS Genet 13:e1006911
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Arens P, van der Sluis T, van’t Westende WPC, Vosman B, Vos CC, Smulders MJM (2007) Genetic population differentiation and connectivity among fragmented Moor frog (Rana arvalis) populations in The Netherlands. Landsc Ecol 22:1489–1500
    Article  Google Scholar 

    Bachmann JC, van Rensburg AJ, Cortazar-Chinarro M, Laurila A, Buskirk JV (2020) Gene Flow Limits Adaptation along Steep Environmental Gradients. The American Naturalist 195:E67–E86
    PubMed  Article  PubMed Central  Google Scholar 

    Balkau BJ, Feldman MW (1973) Selection for migration modification. Genetics 74:171–174
    CAS  PubMed  PubMed Central  Google Scholar 

    Barrett RDH, Schluter D (2008) Adaptation from standing genetic variation. Trends Ecol Evol 23:38–44
    PubMed  Article  PubMed Central  Google Scholar 

    Bates D, Mächler M, Bolker B, Walker S (2015) Fitting linear mixed-effects models using lme4. J Stat Softw 67:1–48.
    Article  Google Scholar 

    Becker CG, Rodriguez D, Longo AV, Talaba AL, Zamudio KR (2012) Disease risk in temperate amphibian populations is higher at closed-canopy sites. PLOS ONE 7:e48205
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Bolnick D, Otto S (2013) The magnitude of local adaptation under genotype-dependent dispersal. Ecol Evol 3:4722–4735
    PubMed  PubMed Central  Article  Google Scholar 

    Bolnick DI, Snowberg LK, Patenia C, Stutz WE, Ingram T, Lau OL (2009) Phenotype-dependent native habitat preference facilitates divergence between parapatric lake and stream stickleback. Evolution 63:2004–2016
    PubMed  Article  PubMed Central  Google Scholar 

    Brown PS, Frye BE (1969) Effects of prolactin and growth hormone on growth and metamorphosis of tadpoles of the frog, Rana pipiens. Gen Comp Endocrinol 13:126–138
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Buxton VL, Ward MP, Sperry JH (2017) Frog breeding pond selection in response to predators and conspecific cues. Ethology 123:397–404
    Article  Google Scholar 

    Campbell-Staton SC, Cheviron ZA, Rochette N, Catchen J, Losos JB, Edwards SV (2017) Winter storms drive rapid phenotypic, regulatory, and genomic shifts in the green anole lizard. Science 357:495–498
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Capblancq T, Luu K, Blum MGB, Bazin E (2018) Evaluation of redundancy analysis to identify signatures of local adaptation. Mol Ecol Resour 18:1223–1233
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Catchen J, Hohenlohe PA, Bassham S, Amores A, Cresko WA (2013) Stacks: an analysis tool set for population genomics. Mol Ecol 22:3124–3140
    PubMed  PubMed Central  Article  Google Scholar 

    Caye K, Jay F, Michel O, François O (2018) Fast inference of individual admixture coefficients using geographic data. Ann Appl Stat 12:586–608
    Article  Google Scholar 

    Clarke RT, Rothery P, Raybould AF (2002) Confidence limits for regression relationships between distance matrices: estimating gene flow with distance. JABES 7:361
    Article  Google Scholar 

    Conesa A, Götz S, García-Gómez JM, Terol J, Talón M, Robles M (2005) Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics 21:3674–3676
    CAS  Article  Google Scholar 

    Delph LF (2018) The study of local adaptation: a thriving field of research. J Heredity 1:1–2
    Article  Google Scholar 

    Denver RJ (1997) Proximate mechanisms of phenotypic plasticity in amphibian metamorphosis. Integr Comp Biol 37:172–184
    CAS  Google Scholar 

    Dyer RJ, Chan DM, Gardiakos VA, Meadows CA (2012) Pollination graphs: Quantifying pollen pool covariance networks and the influence of intervening landscapes on genetic connectivity in the North American understory tree, Cornus florida. Landsc Ecol 27:239–251
    Article  Google Scholar 

    Forester BR, Lasky JR, Wagner HH, Urban DL (2018) Comparing methods for detecting multilocus adaptation with multivariate genotype–environment associations. Mol Ecol 27:2215–2233
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Fox F, Weisberg S (2019) An R Companion to Applied Regression, Third Edition. Sage, Thousand Oaks CA, https://socialsciences.mcmaster.ca/jfox/Books/Companion/

    Fraser DJ, Weir L, Bernatchez L, Hansen MM, Taylor EB (2011) Extent and scale of local adaptation in salmonid fishes: review and meta-analysis. Heredity 106:4040–4420
    Article  Google Scholar 

    Frichot E, François O (2015) LEA: An R package for landscape and ecological association studies. Methods Ecol Evol 6:925–929
    Article  Google Scholar 

    Frichot E, Schoville SD, Bouchard G, François O (2013) Testing for associations between loci and environmental gradients using latent factor mixed models. Mol Biol Evol 30:1687–1699
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Galas L, Raoult E, Tonon M-C, Okada R, Jenks BG, Castaño JP et al. (2009) TRH acts as a multifunctional hypophysiotropic factor in vertebrates. Gen Comp Endocrinol 164:40–50

    Gjertsen AK (2007) Accuracy of forest mapping based on Landsat TM data and a kNN-based method. Remote Sens Environ 110:420–430
    Article  Google Scholar 

    Gosselin T (2018) grur: an R package tailored for RADseq data imputations. R package version 0.0.10. https://github.com/thierrygosselin/grur. https://doi.org/10.5281/zenodo.496176

    Gosselin T, Anderson EC, Bradbury I (2016) assigner: Assignment Analysis with GBS/RAD Data using R. R package version 0.4.1. https://github.com/thierrygosselin/assigner. https://doi.org/10.5281/zenodo.51453

    Gosner KL (1960) A simplified table for staging anuran embryos and larvae with notes on identification. Herpetologica 16:183–190
    Google Scholar 

    Graham LJ, Haines-Young RH, Field R (2017) Metapopulation modeling of long-term urban habitat-loss scenarios. Landsc Ecol 32:989–1003
    PubMed  PubMed Central  Article  Google Scholar 

    Haldane JBS (1930) A mathematical theory of natural and artificial selection. (Part VI, Isolation.). Math Proc Camb Philos Soc 26:220–230
    Article  Google Scholar 

    Hammond SA, Warren RL, Vandervalk BP, Kucuk E, Khan H, Gibb EA et al. (2017) The North American bullfrog draft genome provides insight into hormonal regulation of long noncoding RNA. Nature. Communications 8:1433
    Google Scholar 

    Hangartner S, Laurila A, Räsänen K (2012) Adaptive divergence in Moor Frog (Rana arvalis) populations along an acidification gradient: inferences from Qst–Fst correlations. Evolution 66:867–881
    PubMed  Article  PubMed Central  Google Scholar 

    Hanski I, Mononen T, Ovaskainen O (2011) Eco-evolutionary metapopulation dynamics ant the spatial scale of adaptation. Am Naturalist 177:29–43
    Article  Google Scholar 

    Hedrick PW (2013) Adaptive introgression in animals: examples and comparison to new mutation and standing variation as sources of adaptive variation. Mol Ecol 22:4606–4618
    PubMed  Article  PubMed Central  Google Scholar 

    Hellsten U, Harland RM, Gilchrist MJ, Hendrix D, Jurka J, Kapitonov V et al. (2010) The genome of the western clawed frog Xenopus tropicalis. Science 328:633–636
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Hemmer-Hansen J, Nielsen EE, Therkidsen NO, Taylor MI, Ogden R, Geffen AJ et al. (2013) A genomic island linked to ecotype divergence in Atlantic cod. Mol Ecol 22:2653–2667
    PubMed  Article  PubMed Central  Google Scholar 

    Hendry AP, Day T (2005) Population structure attributable to reproductive time: isolation by time and adaptation by time. Mol Ecol 14:901–916
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Hoban S, Kelley JL, Lottherhos KE, Antolin MF, Bradburd G, Lowry DB et al. (2016) Finding the genomic basis of local adaptation: pitfalls, practical solutions, and future directions. Am Naturalist 188:379–397
    Article  Google Scholar 

    Holt RD, Roy M (2007) Predation Can increase the prevalence of infectious disease. Am Naturalist 169:690–699
    Article  Google Scholar 

    Homola JJ, Loftin CS, Kinnison MT (2019) Landscape genetics reveals unique and shared effects of urbanization for sympatric pool-breeding amphibians. Ecol Evol 9:17799–17823
    Article  Google Scholar 

    Ishwaran H, Kogalur UB (2018) Fast unified random forests for survival, regression, and classification (RF-SRC), R package version 2.7.0

    Ismail SA, Kokko H (2019) An analysis of mating biases in trees. Mol Ecol 29:1–15
    Google Scholar 

    Jenkins DG, Carey M, Czerniewska J, Fletchet J, Hether T, Jones A et al. (2010) A meta-analysis of isolation by distance: relic or reference standard for landscape genetics? Ecography 33:315–320
    Google Scholar 

    Johansson F, Halvarsson P, Mikolajewski DJ, Höglund J (2017) Genetic differentiation in the boreal dragonfly Leucorrhinia dubia in the Palearctic region. Biol J Linn Soc 121:294–304
    Article  Google Scholar 

    Johansson M, Primmer CR, Sahlsten J, Merila J (2005) The influence of landscape structure on occurrence, abundance and genetic diversity of the common frog, Rana temporaria. Global Change Biol 11:1664–1679
    Article  Google Scholar 

    Jombart T (2008) adegenet: a R package for the multivariate analysis of genetic markers. Bioinformatics 24:1403–1405
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Jombart T, Ahmed I (2011) adegenet 1.3-1: new tools for the analysis of genome-wide SNP data. Bioinformatics 27:3070–3071
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Kawecki TJ, Ebert D (2004) Conceptual issues in local adaptation. Ecol Lett 7:1225–1241
    Article  Google Scholar 

    Korhonen L, Korhonen K, Rautiainen M, Stenberg P (2006) Estimation of forest canopy cover: a comparison of field measurement techniques. Silva Fennica 40:577–588
    Article  Google Scholar 

    Legendre L, Legendre P (1998) Numerical ecology. Second English edition. Elsevier Science BV, Amsterdam, The Netherlands

    Legendre P, Oksanen J, ter Braak CJF (2011) Testing the significance of canonical axes in redundancy analysis. Methods Ecol Evol 2:269–277
    Article  Google Scholar 

    Leinonen T, McCairns RJS, O’Hara RB, Merilä J (2013) QST–FST comparisons: evolutionary and ecological insights from genomic heterogeneity. Nat Rev Genet 14:179–190
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Leimu R, Fischer M (2008) A meta-analysis of local adaptation in plants. PLoS ONE 3:e4010
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Lenormand T (2002) Gene flow and the limits to natural selection. Trends Ecol Evol 17:183–189
    Article  Google Scholar 

    Lenhardt PP, Brühl CA, Leeb C, Theissinger K (2017) Amphibian population genetics in agricultural landscapes: does viniculture drive the population structuring of the European common frog (Rana temporaria)? PeerL 5:e3520
    Article  CAS  Google Scholar 

    Lind MI, Johansson F (2007) The degree of adaptive phenotypic plasticity is correlated with the spatial environmental heterogeneity experienced by island populations of Rana temporaria. J Evolut Biol 20:1288–1297
    CAS  Article  Google Scholar 

    Lüdecke D, Makowski D, Waggoner P, Patil I (2020) performance: Assessment of regression models performance. R package version 0.4.6. https://CRAN.R-project.org/package=performance

    Luquet E, Mörch PR, Cortázar‐Chinarro M, Meyer‐Lucht Y, Höglund J, Laurila A (2019) Post‐glacial colonization routes coincide with a life‐history breakpoint along a latitudinal gradient. J Evol Biol 32:356–368
    PubMed  Article  PubMed Central  Google Scholar 

    Luu K, Bazin E, Blum MGB (2017) pcadapt: an R package to perform genome scans for selection based on principal component analysis. Mol Ecol Resour 17:67–77
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Maes GE, Pujolar JM, Hellemans B, Volkaert FA (2006) Evidence for isolation by time in the European eel (Anguilla anguilla). Mol Ecol 15:2095–2107
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Mastretta‐Yanes A, Arrigo N, Alvarez N, Jorgensen TH, Piñero D, Emerson BC (2015) Restriction site-associated DNA sequencing, genotyping error estimation and de novo assembly optimization for population genetic inference. Mol Ecol Resour 15:28–41
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    McRae BH (2006) Isolation by Resistance. Evolution 60:1551–1561
    PubMed  Article  PubMed Central  Google Scholar 

    Meirmans PG (2015) Seven common mistakes in population genetics and how to avoid them. Mol Ecol 24:3223–3231
    PubMed  Article  PubMed Central  Google Scholar 

    Meurling S, Kärvemo S, Chondrelli N, Cortazar-Chinarro M, Åhlén D, Brookes L et al. (2020) Occurrence of Batrachochytrium dendrobatidis in Sweden: higher infection prevalence in southern species. Dis Aquat Organ in press.

    Michel MJ (2011) Spatial dependence of phenotype-environment associations for tadpoles in natural ponds. Evolut Ecol 25:915–932
    Article  Google Scholar 

    Montero-Mendieta S, Tan K, Christmas MJ, Olsson A, Vilà C, Wallberg A et al. (2019) The genomic basis of adaptation to high-altitude habitats in the eastern honey bee (Apis cerana). Mol Ecol 28:746–760.
    PubMed  PubMed Central  Google Scholar 

    Nakajima K, Fujimoto K, Yaoita Y (2005) Programmed cell death during amphibian metamorphosis. Semin Cell Dev Biol 16:271–280
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Nunes AL, Orizaola G, Laurila A, Rebelo R (2014) Rapid evolution of constitutive and inducible defenses against an invasive predator. Ecology 95:1520–1530
    PubMed  Article  PubMed Central  Google Scholar 

    Oksanen J, Guillaume-Blanchet F, Friendly M, Kindt R, Legendre P, McGlinn D et al. (2019) vegan:Community Ecology Package. R package version 2:5–5. https://CRAN.R-project.org/package=vegan

    Palik B, Batzer DP, Buech R, Nichols D, Cease K, Egeland L et al. (2001) Seasonal pond characteristics across a chronosequence of adjacent forest ages in northern Minnesota, USA. Wetlands 21:532–542
    Article  Google Scholar 

    Papaïx J, David O, Lannou C, Monod H (2013) Dynamics of adaptation in spatially heterogeneous metapopulations. PLoS ONE 8:e54697. https://doi.org/10.1371/journal.pone.0054697
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    Paxton RJ, Thorén PA, Tengö J, Estoup A, Pamilo P (1996) Mating structure and nestmate relatedness in a communal bee, Andrena jacobi (Hymenoptera, Andrenidae), using microsatellites. Mol Ecol 5:511–519
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Peterson BK, Weber JN, Kay EH, Fisher HS, Hoekstra HE (2012) Double Digest RADseq: an inexpensive method for de novo SNP discovery and genotyping in model and non-model species. PLOS ONE 7:e37135
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    R Core Team (2019) R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria, https://www.R-project.org/

    Rašić G, Filipović I, Weeks AR, Hoffmann AA (2014) Genome-wide SNPs lead to strong signals of geographic structure and relatedness patterns in the major arbovirus vector, Aedes aegypti. BMC Genomics 15:275
    PubMed  PubMed Central  Article  Google Scholar 

    Reese H, Nilsson M, Pahén TG, Hagner O, Joyce S, Tingelöf U et al. (2003) Countrywide estimates of forest variables using satellite data and field data from the National Forest Inventory. Ambio 32:542–548
    PubMed  Article  PubMed Central  Google Scholar 

    Ribolli J, Hoeinghaus DJ, Johnson JA, Zaniboni-Filho E, de Freitas PD, Galetti PM (2017) Isolation-by-time population structure in potamodromous Dourado Salminus brasiliensis in southern Brazil. Conserv Genet 18:67–76
    Article  Google Scholar 

    Richardson JL, Urban MC, Bolnick DI, Skelly DK (2014) Microgeographic adaptation and the spatial scale of evolution. Trends Ecol Evol 29:165–176
    PubMed  Article  PubMed Central  Google Scholar 

    Richter-Boix A, Katzenberger M, Duarte H, Quintela M, Tejedo M, Laurila A (2015) Local divergence of thermal reaction norms among amphibian populations is affected by pond temperature variation. Evolution 69:2210–2226
    PubMed  Article  PubMed Central  Google Scholar 

    Richter-Boix A, Quintela M, Kierczak M, Franch M, Laurila A (2013) Fine-grained adaptive divergence in an amphibian: genetic basis of phenotypic divergence and the role of nonrandom gene flow in restricting effective migration among wetlands. Mol Ecol 22:1322–1340
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Rödin-Mörch P (2019) Population divergence at different spatial scales in a wide-spread amphibian. PhD thesis. Uppsala University, Sweden
    Google Scholar 

    Rödin-Mörch P, Luquet E, Meyer-Lucht Y, Richter‐Boix A, Höglund J, Laurila A (2019) Latitudinal divergence in a widespread amphibian: contrasting patterns of neutral and adaptive genomic variation. Mol Ecol 28:2996–3011
    PubMed  Article  PubMed Central  Google Scholar 

    Rollins-Smith LA (2009) The role of amphibian antimicrobial peptides in protection of amphibians from pathogens linked to global amphibian declines. Biochim et Biophys Acta 1788:1593–1599
    CAS  Article  Google Scholar 

    Rollins-Smith LA, Conlon JM (2005) Antimicrobial peptide defenses against chytridiomycosis, an emerging infectious disease of amphibian populations. Dev Comp Immunol 29:589–598
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Row JR, Knick ST, Oyler‐McCance SJ, Lougheed SC, Fedy BC (2017) Developing approaches for linear mixed modeling in landscape genetics through landscape-directed dispersal simulations. Ecol Evol 7:3751–3761
    PubMed  PubMed Central  Article  Google Scholar 

    Rundle HD, Nosil P (2005) Ecological speciation. Ecol Lett 8:336–352
    Article  Google Scholar 

    Safner T, Miaud C, Gaggiotti O, Decout S, Rioux D, Zundel S et al. (2011) Combining demography and genetic analysis to asses the population structure of an amphibian in a human-dominated landscape. Conserv Genet 12:161–173
    Article  Google Scholar 

    Santos H, Rousselet J, Magnoux E, Paiva MR, Branco M, Kerdelhué C (2007) Genetic isolation though time: allochronic differentiation of a phonologically atypical population of the pine processionary moth. Proc R Soc B 274:935–941
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Savolainen O, Lascoux M, Merilä J (2013) Ecological genomics of local adaptation. Nat Rev Genet 14:807–820
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Savolainen O, Pyhäjärvi T, Knürr T (2007) Gene flow and local adaptation in trees. Annu Rev Ecol Evol Syst 38:595–619
    Article  Google Scholar 

    Semlitsch RD (2008) Differentiating migration and dispersal processes for pond-breeding amphibians. J Wildl Manag 72:260–267
    Article  Google Scholar 

    Session AM, Uno Y, Kwon T, Chapman JA, Toyoda A, Takahashi S et al. (2016) Genome evolution in the allotetraploid frog Xenopus laevis. Nature 538:336–343
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Sexton JP, Hangartner SB, Hoffmann AA (2014) Genetic isolation by environment or distance: which pattern of gene flow is most common? Evolution 68:1–15
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Shafer ABA, Wolf JBW (2013) Widespread evidence for incipient ecological speciation: a meta-analysis of isolation-by-ecology. Ecol Lett 16:940–950
    PubMed  Article  PubMed Central  Google Scholar 

    Sillero N, Campos J, Bonardi A, Corti C, Creemers R, Crochet P-A et al. (2014) Updated distribution and biogeography of amphibians and reptiles of Europe. Amphib Reptilia 35:1–31
    Article  Google Scholar 

    Skelly DK (2004) Microgeographic countergradient variation in the wood frog, Rana Sylvatica. Evolution 58:160–165
    PubMed  Article  PubMed Central  Google Scholar 

    Smith MA, Green DM (2005) Dispersal and the metapopulation paradigm in amphibian ecology and conservation: are all amphibian populations metapopulations? Ecography 28:110–128
    Article  Google Scholar 

    Storey JD, Bass AJ, Dabney A, Robinson D (2019) qvalue: Q-value estimation for false discovery rate control. R package version 2.16.0. http://github.com/jdstorey/qvalue

    Storfer A, Murphy MA, Evans JS, Goldberg CS, Robinson S, Spear SF et al. (2007) Putting the ‘landscape’ in landscape genetics. Heredity 98:128–142
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Sun Y-B, Xiong Z-J, Xiang X-Y, Liu S-P, Zhou W-W, Tu X-L et al. (2015) Whole-genome sequence of the Tibetan frog Nanorana parkeri and the comparative evolution of tetrapod genomes. Proc Natl Acad Sci USA 112:E1257–E1262
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Tang F, Ishwaran H (2017) Random forest missing data algorithms. Stat Anal Data Min 10:363–377
    PubMed  PubMed Central  Article  Google Scholar 

    Tigano A, Friesen VL (2016) Genomics of local adaptation with gene flow. Mol Ecol 25:2144–2164
    PubMed  Article  PubMed Central  Google Scholar 

    Van Buskirk J (2014) Incipient habitat race formation in an amphibian. J Evolut Biol 27:585–592
    Article  Google Scholar 

    Van Strien MJ, Holderegger R, van Heck HJ (2015) Isolation-by-distance in landscapes: considerations for landscape genetics. Heredity 114:27–37
    PubMed  Article  PubMed Central  Google Scholar 

    Vasemägi A (2006) The adaptive hypothesis of clinal variation revisited: single-locus clines as a result of spatially restricted gene flow. Genetics 173:2411–2414
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Vendrami DLJ, Telesca L, Weigand H, Weiss M, Fawcett K, Lehman K et al. (2017) RAD sequencing resolves fine-scale population structure in a benthic invertebrate: implications for understanding phenotypic plasticity. R Soc Open Sci 4:160548
    PubMed  PubMed Central  Article  Google Scholar 

    Vincent B, Dionne M, Kent MP, Lien S, Bernatchez L (2013) Landscape genomics in atlantic salmon (salmo salar): searching for gene–environment interactions driving local adaptation. Evolution 67:3469–3487
    PubMed  Article  PubMed Central  Google Scholar 

    Vinogradov AE (1998) Genome size and GC-percent in vertebrates as determined by flow cytometry: the triangular relationship. Cytometry 31:100–109
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Vos CC, Jong AGA-D, Goedhart PW, Smulders MJM (2001) Genetic similarity as a measure for connectivity between fragmented populations of the moor frog (Rana arvalis). Heredity 86:598–608
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Wang IJ, Bradburd GS (2014) Isolation by environment. Mol Ecol 23:5649–5662
    PubMed  Article  PubMed Central  Google Scholar 

    Watanabe K, Kazama S, Omura T, Monaghan MT (2014) Adaptive genetic divergence along narrow environmental gradients in four stream insects. PLoS ONE 9:e93055
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Weir BS, Cockerham CC (1984) Estimating F-statistics for the analysis of population structure. Evolution 38:1358–1370
    CAS  Google Scholar 

    Whelan NV, Galaska MP, Sipley BN, Weber JM, Johnson PD, Halanych KM et al. (2019) Riverscape genetic variation, migration patterns, and morphological variation of the threatened Round Rocksnail, Leptoxis ampla. Mol Ecol 28:1593–1610
    PubMed  Article  PubMed Central  Google Scholar 

    Whitlock MC, McCauley DE (1999) Indirect measures of gene flow and migration: FST ≠ 1/(4 Nm +1). Heredity 82:117
    PubMed  Article  PubMed Central  Google Scholar 

    Wright S (1943) Isolation by distance. Genetics 28:114–138
    CAS  PubMed  PubMed Central  Google Scholar 

    Yandell M, Ence D (2012) A beginner’s guide to eukaryotic genome annotation. Nat Rev Genet 13:329–342
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Yeaman S, Otto SP (2011) Establishment and maintenance of adaptive genetic divergence under migration, selection, and drift. Evolution 65:2123–2129
    PubMed  Article  PubMed Central  Google Scholar 

    Youngquist M, Inoue K, Berg D, Boone MD (2017) Effects of land use on population presence and genetic structure of an amphibian in an agricultural landscape. Landsc Ecol 32:147–162
    Article  Google Scholar 

    Yu L, Wang G-D, Ruan J, Chen Y-B, Yang C-P, Cao X et al. (2016) Genomic analysis of snub-nosed monkeys (Rhinopithecus) identifies genes and processes related to high-altitude adaptation. Nat Genet 48:947–952
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Zamudio KR, Wieczorek AM (2007) Fine-scale spatial genetic structure and dispersal among spotted salamander (Ambystoma maculatum) breeding populations. Mol Ecol 16:257–274
    PubMed  Article  PubMed Central  Google Scholar 

    Zellmer AJ (2018) Microgeographic morphological variation across larval wood frog populations associated with environment despite gene flow. Ecol Evol 8:2504–2517
    PubMed  PubMed Central  Article  Google Scholar 

    Zhang H, Xu Q, Krajewski S, Krajewska M, Xie Z, Fuess S et al. (2000) BAR: An apoptosis regulator at the intersection of caspases and Bcl-2 family proteins. Proc Natl Acad Sci USA 97:2597–2602
    CAS  PubMed  Article  PubMed Central  Google Scholar  More

  • in

    Macroecological laws describe variation and diversity in microbial communities

    Data
    All the data sets analyzed in this work have been previously published and were obtained from EBI Metagenomics69. Previous publications (Supplementart Table 1) report the original experiments and corresponding analysis. In order to test the robustness of the macroecological laws and the modeling framework presented in this work, we considered 7 data sets that differ not only for the biome considered, but also for the sequencing techniques and the pipeline used to process the data. Data sets were selected to represent a wide set of biomes. We considered only data sets with at least 50 samples with more than 104 reads. No data set was excluded a-posteriori.
    Sampling and compositional data
    In order to study how (relative) abundance varies across communities and species, one needs to remove the effect of sampling noise, as it is not a biologically informative source of variation. By explicitly modeling sampling (Supplementary Note 2), one finds that the probability of observing n reads of species i in a sample with N total number of reads, is given by

    $${P}_{i}(n| N)=int_{0}^{1}dx,{rho }_{i}(x),left(begin{array}{*{20}{c}} {n} \ {N} end{array}right){x}^{n}{(1-x)}^{N-n},,$$
    (6)

    where ρi(x) is the AFD, i.e., the probability (over communities or times) that the relative abundance of i is equal to x. Note that this equation does not assume anything about independence across species or communities. It only assumes the sampling process is carried independently across communities.
    Since the random variable xi, whose distribution is ρi(x), is a relative abundance, one has that ∑ixi = 1 (i.e., the data are compositional70). As discussed in Supplementary Note 2, given the range of variation of the empirical relative abundances, one can substitute Eq. (6) with

    $${P}_{i}(n| N)=int_{0}^{infty }dx,{rho }_{i}(x)frac{{(xN)}^{n}}{n!}{e}^{-xN},,$$
    (7)

    and the condition (sum)ixi = 1 to ({sum }_{i}{bar{x}}_{i}=1), where ({bar{x}}_{i}=mathop{int}nolimits_{0}^{infty }dx,{rho }_{i}(x)x) is the mean value of xi. Under this assumption, one can also take the limits of the integration from 0 to ∞, instead of considering them from 0 to 1, as the contribution of the integrand from 1 to ∞ is negligible.
    Note that, because of sampling, the average of a function f(x) over the pdf ρ(x) differs in general from the average of f(n/N) over P(n∣N)

    $$int_{0}^{1}dx,rho (x)f(x), ne, sum _{n = 0}^{N}P(n| N)fleft(frac{n}{N}right)=int_{0}^{1}dx,rho (x)sum _{n = 0}^{N}fleft(frac{n}{N}right)frac{{(xN)}^{n}}{n!}{e}^{-xN},,$$
    (8)

    and the inequality becomes equality only if f(x) is linear. The important difference between the right- and the left-hand side is often neglected in the literature. In fact, the right-hand side is a good approximation of the left-hand size only in the limit xN ≫ 1, which is far from being realized in the data for most of the species. Supplementary Note 2 introduces a method to reconstruct the moments of ρ(x) from the moments of P(n∣N). More generally, I show that it is possible to infer the moment generating function of ρ(x) from the data, which allow to reconstruct the shape of the empirical ρ(x).
    Excluding competitive exclusion
    A Gamma-distributed AFD implies that all the species present in a community of a biome are present in all the communities from that biome. Therefore, when a species is not observed is because it is undetected due to sampling errors. I test this claim in two different ways. First, it is shown that one can in fact predict the occupancy of a species from its abundance fluctuations. Secondly, I show that a model without true absences is statistically more supported than a model where species are allowed to be absent.
    The first way to test this hypothesis is to directly test its immediate prediction: if the absence is a consequence of sampling, one should be able to predict occupancy of a species (the probability that a species is present) simply from its average and variance of abundance (together with the total number of reads of each sample). In particular, assuming a Gamma AFD, the occupancy of species i is given by

    $$langle {o}_{i}rangle =1-frac{1}{T}sum _{s}P(0| {N}_{s})=1-frac{1}{T}sum _{s = 1}^{T}{left(1+frac{{bar{x}}_{i}{N}_{s}}{{beta }_{i}}right)}^{-{beta }_{i}},,$$
    (9)

    where Ns is the total number of reads in sample s, T is the total number of samples, and ({beta }_{i}={bar{x}}_{i}^{2}/{sigma }_{{x}_{i}}^{2}). As shown in Fig. 2 and in Supplementary Fig. 3, this prediction well reproduces the observed occupancy across species. The prediction of Eq. (12) also matches the occupancy of temporal (longitudinal) data Supplementary Fig. 20.
    The second, more rigorous, way to test the hypothesis that (most) species are always present is to use model selection. In this context we want to compare two (or more) models that aim at describing the observed number of reads of each species starting from alternative hypothesis. In particular I compare a purely Gamma AFD with a zero-inflated Gamma, which reads

    $${varrho }_{i}(x| {vartheta }_{i},{beta }_{i},{bar{x}}_{i})={vartheta }_{i}delta (x)+(1-{vartheta }_{i})frac{1}{Gamma ({beta }_{i})}{left(frac{{beta }_{i}}{{bar{x}}_{i}}right)}^{{beta }_{i}}{x}^{{beta }_{i}-1}exp left(-{beta }_{i}frac{x}{{bar{x}}_{i}}right),,$$
    (10)

    where ϑi is the probability that a species is truly absent in a community and δ( ⋅ ) is the Dirac delta distribution. Our goal is to test whether the ϑis are significantly different from zero. Since the two models are nested, one can compare the maximum likelihood estimator in the case ϑi = 0 with the (maximum) likelihood marginalized over ϑ (which has prior μ(ϑ)). Given the number of reads ({n}_{i}^{s}) of species i in community s, with Ns total number of reads, one can compute the ratio (Supplementary Note 4)

    $${ell }_{i}=frac{mathop{max }nolimits_{bar{x},beta }{prod }_{s}int dx{varrho }_{i}(x| 0,beta ,bar{x})frac{{(x{N}_{s})}^{{n}_{i}^{s}}}{{n}_{i}^{s}!}{e}^{-x{N}_{s}}}{int dvartheta ,mu (vartheta )left(mathop{max }nolimits_{bar{x},beta }{prod }_{s}int dx{varrho }_{i}(x| vartheta ,beta ,bar{x})frac{{(x{N}_{s})}^{{n}_{i}^{s}}}{{n}_{i}^{s}!}{e}^{-x{N}_{s}}right)},,$$
    (11)

    where μ(ϑ) is a prior over ϑ. If ℓi  > 1, the model with ϑi = 0 is more strongly supported than the model with ϑ ≠ 0. Under Beta prior with parameters 0.25 and 8, one obtains that ℓi  > 1 in 98.8% of the cases (averaged across data sets, ranging from 94.4 to 99.7%) and ℓi  > 100 in 97.5% cases (ranging from 92.8 to 99.2%). See Supplementary Note 4 for a more detailed description of the methodology and Supplementary Fig. 6 for results obtained with other priors.
    Prediction of macroecological patterns
    Given laws #1, #2, and #3, the probability to observe n reads of a randomly chosen species in a sample with N total reads is

    $$P(n| N)=int_{-infty }^{infty }deta ,frac{Gamma (beta +n)}{n!Gamma (beta )}{left(frac{{e}^{eta }N}{beta +{e}^{eta }N}right)}^{n}{left(frac{beta }{beta +{e}^{eta }N}right)}^{beta }frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}},,$$
    (12)

    where (eta ={mathrm{log}},(bar{x})). All the properties of species are fully specified by its mean abundance (bar{x}={e}^{eta }). The probability of observing n reads of species with average abundance (bar{x}) in a sample with N total number of reads is therefore

    $$P(n| N,bar{x})=frac{Gamma (beta +n)}{n!Gamma (beta )}{left(frac{bar{x}N}{beta +bar{x}N}right)}^{n}{left(frac{beta }{beta +bar{x}N}right)}^{beta },.$$
    (13)

    The predictions for the patterns shown in Fig. 3 are reported here. The full derivation of this and other patterns is presented in Supplementary Note 8.
    The total number of observed species in a sample with N total number of reads can be easily calculated using Eq. (12). The probability of not observing a species is simply P(0∣N). The expected number of distinct species 〈s(N)〉 in a sample with N reads is therefore

    $$langle s(N)rangle ={s}_{tot}left(1-P(0| N)right)={s}_{tot}left(1-int_{-infty }^{infty }deta ,frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}},{left(frac{beta }{beta +{e}^{eta }N}right)}^{beta }right),,$$
    (14)

    where stot is the total number of species in the biome (including unobserved ones, see Supplementary Note 7). Note that stot is (substantially) larger than sobs, the number of different species observed in the union of all the communities, which can instead be written as

    $$langle {s}_{obs}rangle ={s}_{tot}left(1-int_{-infty }^{infty }deta ,frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}},{left(prod _{s = 1}^{T}frac{beta }{beta +{e}^{eta }{N}_{s}}right)}^{beta }right),.$$
    (15)

    Figure 3a shows that the prediction of Eq. (14) correctly matches the data (Supplementary Fig. 13).
    The SAD, one of the most studied patterns in ecology and directly related to the Relative Species Abundance35, is defined as the fraction of species with a given abundance. According to our model, the expected SAD is given by

    $$langle {Phi }_{n}(N)rangle := frac{langle {s}_{n}(N)rangle }{langle s(N)rangle }=frac{P(n| N)}{1-P(0,N)}=frac{int_{-infty }^{infty }deta ,frac{Gamma (beta +n)}{n!Gamma (beta )}{left(frac{{e}^{eta }N}{beta +{e}^{eta }N}right)}^{n}{left(frac{beta }{beta +{e}^{eta }N}right)}^{beta }frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}}}{1-int_{-infty }^{infty }deta ,{left(frac{beta }{beta +{e}^{eta }N}right)}^{beta },frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}}},,$$
    (16)

    where 〈sn(N)〉 is the number of species with n reads in a sample with N total number of reads. The cumulative SAD is defined as

    $$langle {Phi }_{n}^{ ,{ > },}(N)rangle := sum _{m = n}^{infty }langle {Phi }_{m}(N)rangle =frac{int mathop{int}nolimits_{-infty }^{infty },{I}_{frac{{e}^{eta }N}{beta +{e}^{eta }N}}(n,beta )frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}}}{1-mathop{int}nolimits_{-infty }^{infty }eta ,{left(frac{beta }{beta +{e}^{eta }N}right)}^{beta },frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}}},,$$
    (17)

    where Ip(n, β) is the regularized incomplete Beta function. Figure 3b shows that the Eq. (17) captures the empirical cumulative SAD (Supplementary Fig. 17).
    The occupancy probability is defined as the probability that a species is present in a given fraction of communities. This quantity has been extensively studied in a variety of contexts (from genomics71 to Lego sets and texts72) and has been more recently considered in microbial ecology37. The three macroecological laws predict (see derivation in Supplementary Note 8)

    $${p}_{obs}(o)=frac{mathop{int}nolimits_{-infty }^{infty }deta ,sum_{t = 1}^{T}delta left(o-1+frac{1}{T}mathop{sum }nolimits_{s = 1}^{T}{left(frac{beta }{beta +{e}^{eta }{N}_{s}}right)}^{beta }right)frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}}mathop{prod }nolimits_{s = 1}^{T}left(1-{left(frac{beta }{beta +{e}^{eta }{N}_{s}}right)}^{beta }right)}{mathop{int}nolimits_{-infty }^{infty }deta ,frac{exp left(-frac{{(eta -mu )}^{2}}{2{sigma }^{2}}right)}{sqrt{2pi {sigma }^{2}}},mathop{prod }nolimits_{s = 1}^{T}left(1-{left(frac{beta }{beta +{e}^{eta }{N}_{s}}right)}^{beta }right)},,$$
    (18)

    where δ( ⋅ ) is a Dirac delta function. Figure 3c compares the prediction of Eq. (18) with the data (Supplementary Fig. 15).
    Occupancy (the fraction of communities where a species is found) and abundance are not independent properties, and their relative dependence is often referred to as occupancy-abundance relationship21 Given an average (relative) abundance (bar{x}=exp (eta )), the expected occurrence is

    $${langle orangle }_{eta }=1-frac{1}{T}sum _{s = 1}^{T}P(0| {N}_{s},bar{x})=1-frac{1}{T}sum _{s = 1}^{T}{left(frac{beta }{beta +bar{x}{N}_{s}}right)}^{beta },,$$
    (19)

    Figure 3d shows the comparison between data and predictions (Supplementary Fig. 16). These predictions are also tested for temporal (longitudinal) data in Supplementary Figs. 22–24.
    Transition probabilities in longitudinal data
    For longitudinal data, in addition to the stationary AFD, one can study the probability ({rho }_{i}(x^{prime} ,t+Delta t| x,t)) that a species i has abundance (x^{prime}) at time t + Δt conditioned on having abundance x at time t. Instead of focusing on the full distribution, we study its first two (conditional) central moments, i.e. the average and variance of the abundance at t + Δt conditioned to abundance x at time t. In the analysis of the data stationarity is assumed (the distribution ({rho }_{i}(x^{prime} ,t+Delta t| x,t)) depends on Δt but not on t). I also assume that the dynamics of different species are governed by similar equations that only differ in their parameters. One would like therefore to average over species, by properly rescaling their abundances. The average over species is potentially problematic, as it could add a spurious effect to the conditional averages. For instance, only species with larger fluctuations would appear for extreme values of the initial abundance. In order to avoid these problems, instead of consider the actual abundance, its cumulative probability distribution value (calculated using the empirical AFD of each species) was used, that is referred as “quantile abundance”. This is equivalent to rank the abundances of each species over communities and use the (relative) ranking of each community instead of the abundance. A value equal to 0 corresponds to the lowest observed abundance, and a value equal to 1 to the highest. By definition, the quantile abundance is always uniformly distributed.
    Ruling out demographic stochasticity
    Demographic stochasticity can reproduce a Gamma AFD. A birth, death, and immigration process has a Gamma as stationary distribution35. In the limit of large populations sizes, it corresponds to the following equation35

    $$frac{dx}{dt}=m-(d-b)x+sqrt{(b+d)x}xi (t),,$$
    (20)

    where m is the migration rate, while b and d are the per-capita birth and death rate. The Gaussian white noise term ξ(t) has mean zero and time-correlation (langle xi (t)xi (t^{prime} )rangle =delta (t-t^{prime} )). The stationary distribution of this process turns out to be

    $$rho (x)=frac{1}{Gamma left(2frac{m}{b+d}right)}{left(frac{b+d}{2(d-b)}right)}^{-2frac{m}{b+d}}{x}^{2frac{m}{b+d}-1}exp left(-2frac{d-b}{b+d}xright),.$$
    (21)

    The average abundance is equal to (bar{x}=m/(d-b)), while the variance turns out to be ({sigma }_{x}^{2}=(m/2)(b+d)/{(b-d)}^{2}). The square of the coefficient of variation would therefore be equal to (b + d)/(2m).
    More generally, one can assume that all the parameters are species dependent, and the population of species i is described by

    $$frac{d{x}_{i}}{dt}={m}_{i}-({d}_{i}-{b}_{i}){x}_{i}+sqrt{({b}_{i}+{d}_{i}){x}_{i}}{xi }_{i}(t),,$$
    (22)

    where (langle {xi }_{i}(t){xi }_{j}(t^{prime} )rangle ={delta }_{ij}delta (t-t^{prime} )) was assumed.
    Taylor’s Law and the wide variation of average abundance together imply that mi/(bi + di) is constant while mi/(di − bi) varies across species on several orders of magnitudes. This imposes a constraint on the variation of parameter values across species.
    For instance, one can consider the scenario where species migrate to local communities from a common species pool (metacommunity). As abundance in the metacommunity varies across species the migration rate is a species-dependent quantity. Under neutrality, the per-capita birth and death rates in the local communities are constant and independent of the identity of the species. In this case mi depends on the species, while b and d do not. One could recover the Lognormal MAD by imposing that mi is Lognormally distributed. On the other hand, this model would fail in reproducing Taylor’s law with exponent 2, as it would predict and exponent 1.
    More in general, the condition imposed on the parameters corresponds to an unnatural fine-tuned relationship between migration, birth, and death rates. Variation of the average abundance is observed across, at least, 7 orders of magnitudes. In order to reproduce this variation across species and Taylor’s law with exponent 2, the range of variability of (bi − di)/(bi +  di) should be of the same order. It is unrealistic that the relative difference between birth and death rates, which have strong and direct connection to fundamental biological processes, vary so much across bacterial species. It is important to underline however, that the model of Eq. (22) can, in fact, for a proper parameterization, explain the observed variation of the data. But the choice of parameters explaining the empirical variation require for achieving this goal requires careful and unrealistic fine-tuning of the microscopic parameters.
    Stochastic logistic model
    The SLM is defined as

    $$frac{d{x}_{i}}{dt}=frac{{x}_{i}}{{tau }_{i}}left(1-frac{{x}_{i}}{{K}_{i}}right)+sqrt{frac{{sigma }_{i}}{{tau }_{i}}}{x}_{i}{xi }_{i}(t),,$$
    (23)

    where ξ(t) is a Gaussian white noise term with mean zero and correlation (langle {xi }_{i}(t){xi }_{j}(t^{prime} )rangle ={delta }_{ij}delta (t-t^{prime} )). Taylor’s Law and the observed Lognormal MAD constraints the parameter value. The parameters 1/τi, Ki and σi are the intrinsic growth rate, the carrying capacity and the coefficient of variation of the growth-rate fluctuations. Taylor’s Law requires σi = σ (independently of i). Since the average abundance of the SLM is ({bar{x}}_{i}={K}_{i}(1-{sigma }_{i}/2)), if σi = σ, the average abundance and the carrying capacity turn out to be proportional to each other. The lognormal MAD implies therefore that the Kis are lognormally distributed. The stationary distribution corresponding to Eq. (23) reads

    $${rho }_{i}(x)=frac{1}{Gamma (2{sigma }_{i}^{-1}-1)}{left(frac{2}{{K}_{i}{sigma }_{i}}right)}^{2{sigma }_{i}^{-1}-1}exp left(-frac{2}{{K}_{i}{sigma }_{i}}xright){x}^{2{sigma }_{i}^{-1}-2},.$$
    (24)

    The parameter τi does not affect stationary properties, but determines the timescale of relaxation to the stationary distribution. For small deviation of abundance from the average and for large times, the conditional expected abundance behaves as

    $${langle {x}_{i}(t+Delta t)rangle }_{{x}_{i}(t)}={bar{x}}_{i}+left({x}_{i}(t)-{bar{x}}_{i}right){e}^{-frac{Delta t}{{tau }_{i}}},.$$
    (25)

    From the slopes of Fig. 4g one can then determine the timescales τi, which turn out to be approximately equal to 19 h. In Fig. 4 it was assumed τi = 19 h for all species.
    Equation (23) can emerge as effective description of more complicated coupled equations. For instance, it is possible to show that a Lotka-Volterra system of equation with random interactions reduces to Eq. (23) (with colored noise to be self-consistently determined)42. If the coefficient of variation of the interaction coefficient does not increase with the number of species (e.g., if it is constant) then the Lotka-Volterra equations can be effectively approximated with Eq. (23).
    The noise term in Eq. (23) can be interpreted as corresponding to environmental fluctuations. These fluctuations are typically known to have a characteristic timescale and are not white40,41. Supplementary Note 13 and Supplementary Fig. 25 show that colored noise in Eq. (23) does not affect significantly the predictions obtained with the SLM with white noise.
    Reporting summary
    Further information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    A multi-scale eco-evolutionary model of cooperation reveals how microbial adaptation influences soil decomposition

    Construction of the five-compartment model
    Here we explain the construction of the five-compartment model (Fig. 1a). This is step 1 among the four steps described in “Results” section (subsection “Ecosystem dynamics at microsite scale”). We use upper bars in our initial notations to indicate parameters prior to rescaling.
    The five-compartment model captures the stochastic processes acting at the level of C, D, M, Z, X entities (molecules, cells) (Fig. 1a) within a microsite. Dynamics of C, D, M, Z, X occur in continuous time. Mt is the number of cells at time t. Ct, Dt, Zt are the numbers of SOC molecules, DOC molecules, and enzyme molecules respectively. Xt is the number of complexes formed by an enzyme molecule binding a SOC molecule. There are constant external sources of SOC and DOC. When a cell dies, a fraction p of the molecules released are recycled into SOC, while the rest is recycled into DOC. A fraction l of dead microbes and deactivated enzymes may be lost due to leaching.
    We denote by α the structural cost of a cell, which is the equivalent in number of DOC molecules of one cell (without storage). We denote by (alpha ^{prime}) the energetic cost of a cell, which is the number of DOC molecules consumed to produce the energy needed for the synthesis reactions involved in the production of one cell. We denote by β the equivalent in number of DOC molecules of one SOC molecule, and the structural and energetic cost of producing one molecule of enzyme by ρ and (rho ^{prime}), respectively. We assume that the energetic costs are carbon released by cells as CO2 (cell respiration) that diffuses out of the system instantly. We define the biomass production fraction and enzyme allocation fraction as

    $${bar{gamma }}_{M}:=frac{1}{alpha +alpha ^{prime} },quad {bar{gamma }}_{Z}:=frac{1}{rho +rho ^{prime} }.$$
    (1)

    The event times are given by independent exponential random variables whose parameters are defined by event rates (Supplementary Tables 2–4). These event rates give an approximation of the average frequency of each event. The rates of cell growth and enzyme production depend on the trait φ. Once a cell has doubled its initial size, reproduction occurs by releasing the mother cell at its initial size, and the daughter cell at its same size. The cell must therefore take up and store its structural and energetic cost, ((alpha +alpha ^{prime} )), in DOC molecules in order to reproduce. We denote N the number of uptake events before reproduction. The number of DOC molecules taken up at each uptake event is then ((alpha +alpha ^{prime} )/N), hence the notation ({{bf{1}}}_{{Dge (alpha +alpha ^{prime} )/N}}) which equals 1 if (Dge (alpha +alpha ^{prime} )/N), and 0 otherwise. The same notation, ({{bf{1}}}_{{Dge rho +rho ^{prime} }}), is used for the production event of an enzyme molecule. Uptake is stochastic, but reproduction is deterministic, which means that when a cell has performed N uptake events, it reproduces with probability 1. A larger N means a larger number of uptake events between 2 reproduction events, which also means less DOC molecules taken up at each uptake event. The model tracks the dynamics of the number of cells, SOC, DOC, enzyme molecules, and also of the DOC stored in each cell.
    Enzyme–substrate complexes form at rate ({bar{lambda }}^{k}) as one enzyme molecule (e.g. cellulase) bind one SOC molecule (e.g. cellulose). A complex may either dissociate (with no decomposition) at rate ({bar{lambda }}_{-1}^{varepsilon }), or react at rate ({bar{mu }}^{varepsilon }) and convert the molecule of SOC into β molecule of DOC while the enzyme is released and free again to react with new molecules of SOC (Supplementary Table 3).
    System size k does not appear in this system of equations, yet it enters the volume-dependent parameters of the model, IC, ID, θ, and Km. We denote V the unit volume of soil that contains on average one microbial cell and the corresponding equilibrium of carbon mass of SOC, DOC and enzymes. The system size k is the number of well-mixed unit volumes in one microsite, which determines the number of cells sharing SOC, DOC, and enzymes. The volume of one microsite is therefore k × V. Increasing k amounts to increasing microsite volume, the number of cells sharing resources in one microsite, the amount of resources per microsite, and the volume-dependent parameters, such as the amount of SOC entering a microsite per unit of time, IC. In our analysis, the unit volume V is fixed, and we vary k to investigate the effect of microsite volume on the system’s eco-evolutionary dynamics. With very large k, the hybrid model can be approximated by a fully deterministic model which takes the form of a system of four ordinary differential equations (see Supplementary Information 3.3 and Supplementary Fig. 1), similar to the microbial decomposition model first introduced by Allison et al.53. However, empirical data suggest that k is of the order of 10–10054. When k = 1, there is only one cell in the microsite, which volume is V defined as the unit soil volume expected to contain a single cell. A value of k greater than 1 means that each microsite contains k cells and k times the amount of SOC, DOC and enzyme molecules of 1 cell; thus, the microsite volume is k × V, and volume-dependent parameters are rescaled by k. Specifically, there are four volume-dependent parameters: the external input of C, ({bar{I}}_{C}^{k}), the external input of D, ({bar{I}}_{D}^{k}), the half-saturation constant of DOC uptake, ({bar{K}}_{m}^{k}), and the encounter intensity of two given SOC and enzyme molecules, ({bar{lambda }}^{k}). The external inputs increase proportionally with the volume of the microsite, while the encounter intensity of two given molecules in a microsite decreases as its volume increases. The half-saturation is inversely proportional to the affinity between a given cell M and a given DOC molecule d, which decreases with increasing microsite volume. We thus obtain the following scaling relationships:

    $${bar{I}}_{C}^{k}=k{bar{I}}_{C}, {bar{I}}_{D}^{k}=k{bar{I}}_{D}, {bar{lambda }}^{k}=frac{lambda }{k} {rm{and}} {bar{K}}_{m}^{k}=k{bar{K}}_{m}.$$
    (2)

    In our simulations, we generally assume that k is equal to 10, to match the empirical observation that (cells) in soil habitat tend to interact with 10 to 100 other cells at all time54.
    Derivation of the hybrid model
    In Supplementary Information 3, we present the next two steps (2 and 3 in “Ecosystem dynamics at microsite scale” of “Results” section) to derive the hybrid model on which all our results are based. In Supplementary Information 3.1, we explain how the dynamics generated by the five-compartment model can be captured with a reduced model with four-state variables (step 2). In Supplementary Information 3.2, we explain how the stochastic-deterministic PDMP model can be derived from the stochastic four-state variable model (step 3). In the hybrid model (step 4), only cell death remains stochastic, and cell dynamics is measured in unit of number of individuals (M), while other entities are now in carbon mass unit. The rescaled SOC, DOC and enzyme abundances are denoted with lower case letters c, d, and z.
    Simulation algorithm for the hybrid spatial model
    One technical benefit of the hybrid model is its much greater computational tractability. Here we describe the algorithm used to perform simulations of the hybrid model. We ran the model on single microsite to produce the simulations reported in Fig. 2. We ran the model on a 10 × 10 lattice of microsites for the subsequent figures. The algorithm is based on the Gillepsie algorithm55 as used in Champagnat et al.37, Fournier and Méléard56, which straightforwardly extends to the simulation of PDMPs.
    To couple PDMP models across microsites, we account for the DOC and dispersal of cells between adjacent microsites. The DOC diffusion between microsites is modeled by approximating a continuous diffusion with a Euler scheme in which time is discretized with a fixed time step interval, τdiff. τdiff is chosen sufficiently small to provide a fine enough discretization of the DOC diffusion.
    A simulation starts with a given amount of M, z, c, and d in each microsite at time t = 0, while the initial amount of DOC stored within each cell is determined uniformly at random. Two stochastic events (death of a cell) may not occur at the same time. Assume that the process has been computed until time ti; to continue the computation to time ti+1, we proceed as follows.
    First, we simulate T, an exponential random variable with parameter r(ti) = dMM(ti), which corresponds to the death rate of the total cell population at time ti (M(ti) being the total number of cells on the entire lattice). We then compute

    $${t}_{i+1}:={t}_{i}+min left(T,{tau }_{{rm{diff}}}right).$$

    To obtain c(ti+1), d(ti+1), and z(ti+1) in each microsite at time ti+1, and the variation in amount of DOC stored within a cell in the corresponding microsite, we use a Euler scheme that solves the dynamical system

    $$left{begin{array}{l}hskip -136pt dot{c}(t)={I}_{C}-{l}_{C}c-theta zc,\ dot{d}(t)={I}_{D}-{l}_{D}d+theta zc+(1-l){d}_{Z}z-{V}_{max }frac{d}{{K}_{m}{,}+{,}d}{omega }_{M}M,\ hskip -84pt dot{z}(t)=varphi {gamma }_{Z}{V}_{max }frac{d}{{K}_{m}{,}+{,}d}{omega }_{M}M-{d}_{Z}z,\ hskip -93pt dot{Delta }(t)=(1-varphi ){gamma }_{M}{V}_{max }frac{d}{{K}_{m}{,}+{,}d}{omega }_{M},end{array}right.$$

    in each microsite between ti and ti+1, where M is the number of cells in the microsite at time ti, Δ gives the amount variation of DOC stored within a cell, Δ(ti) = 0 and the other initial conditions are the biomass of c, d, z in the corresponding microsite at time ti.
    Note that, within a microsite, the variation of stored DOC is the same for all cells and corresponds to Δ(ti+1). Hence, this amount is added to the amount of DOC stored within each cell living in the corresponding microsite. If, for a cell j, the resulting amount Sj(Ti) + Δ(ti+1) is over ωM, a new cell appears. The amount of stored DOC within the new cell and the mother cell is then updated: Sj(ti+1) = (Sj(Ti) + Δ(ti+1) − ωM)/2. Otherwise, Sj(ti+1) = Sj(Ti) + Δ(ti+1). To determine the position of the new cell, the following steps are taken:

    A uniform random variable ϑ1 in [0, 1] is simulated.

    If ϑ1  More

  • in

    Estimating retention benchmarks for salvage logging to protect biodiversity

    1.
    Seidl, R., Schelhaas, M.-J., Rammer, W. & Verkerk, P. J. Increasing forest disturbances in Europe and their impact on carbon storage. Nat. Clim. Chang. 4, 806–810 (2014).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 
    2.
    Kurz, W. et al. Mountain pine beetle and forest carbon feedback to climate change. Nature 452, 987–990 (2008).
    ADS  CAS  PubMed  Google Scholar 

    3.
    Turner, M. G. Disturbance and landscape dynamics in a changing world. Ecology 91, 2833–2849 (2010).
    PubMed  Google Scholar 

    4.
    Müller, J. et al. Increasing disturbance demands new policies to conserve intact forest. Conserv. Lett. 12, e12449 (2019).
    Google Scholar 

    5.
    Lindenmayer, D., Burton, P. J. & Franklin, J. F. Salvage Logging and its Ecological Consequences. (Island Press, 2008).

    6.
    Leverkus, A. B. et al. Salvage logging effects on regulating ecosystem services and fuel loads. Front. Ecol. Environ. 18, 391–400 (2020).
    Google Scholar 

    7.
    Thorn, S. et al. Impacts of salvage logging on biodiversity-a meta-analysis. J. Appl. Ecol. 55, 279–289 (2018).
    PubMed  Google Scholar 

    8.
    Cobb, T. P. et al. Effects of postfire salvage logging on deadwood-associated beetles. Conserv. Biol. 25, 94–104 (2011).
    CAS  PubMed  Google Scholar 

    9.
    Leverkus, A. B., Lindenmayer, D. B., Thorn, S. & Gustafsson, L. Salvage logging in the world’s forests: Interactions between natural disturbance and logging need recognition. Glob. Ecol. Biogeogr. 27, 1140–1154 (2018).
    Google Scholar 

    10.
    Morissette, J. L., Cobb, T. P., Brigham, R. M. & James, P. C. The response of boreal forest songbird communities to fire and post-fire harvesting. Can. J. Res. Can. Rech. 32, 2169–2183 (2002).
    Google Scholar 

    11.
    Georgiev, K. B. et al. Salvage logging changes the taxonomic, phylogenetic and functional successional trajectories of forest bird communities. J. Appl. Ecol. 1365-2664, 13599 (2020).
    Google Scholar 

    12.
    Lindenmayer, D. B., Mcburney, L., Blair, D., Wood, J. & Banks, S. C. From unburnt to salvage logged: quantifying bird responses to different levels of disturbance severity. J. Appl. Ecol. 55, 1626–1636 (2018).

    13.
    Blair, D. P., McBurney, L. M., Blanchard, W., Banks, S. C. & Lindenmayer, D. B. Disturbance gradient shows logging affects plant functional groups more than fire. Ecol. Appl. 26, 2280–2301 (2016).
    PubMed  Google Scholar 

    14.
    Noss, R. F. & Lindenmayer, D. B. The ecological effects of salvage logging after natural disturbance. Conserv. Biol. 20, 946–948 (2006).
    Google Scholar 

    15.
    Hutto, R. L. Toward meaningful snag-management guidelines for postfire salvage logging in North American conifer forests. Conserv. Biol. 20, 984–993 (2006).
    PubMed  Google Scholar 

    16.
    Hutto, R. L. The ecological importance of severe wildfires: some like it hot. Ecol. Appl. 18, 1827–1834 (2008).
    PubMed  Google Scholar 

    17.
    Thorn, S., Müller, J. & Leverkus, A. B. Preventing European forest diebacks. Science 365, 1388.2–1388 (2019).
    Google Scholar 

    18.
    Stokstad, E. Salvage logging research continues to generate sparks. Science 311, 761 (2006).
    CAS  PubMed  Google Scholar 

    19.
    Franklin, J. F. et al. Threads of continuity: ecosystem disturbances, biological legacies and ecosystem recovery. Conserv. Biol. Pract. 1, 8–16 (2000).
    Google Scholar 

    20.
    Lindenmayer, D., Thorn, S. & Banks, S. Please do not disturb ecosystems further. Nat. Ecol. Evol. 1, 0031 (2017).
    Google Scholar 

    21.
    Burivalova, Z., Şekercioğlu, Ç. H. & Koh, L. P. Thresholds of logging intensity to maintain tropical forest biodiversity. Curr. Biol. 24, 1893–1898 (2014).
    CAS  PubMed  Google Scholar 

    22.
    França, F. M., Frazão, F. S., Korasaki, V., Louzada, J. & Barlow, J. Identifying thresholds of logging intensity on dung beetle communities to improve the sustainable management of Amazonian tropical forests. Biol. Conserv. 216, 115–122 (2017).
    Google Scholar 

    23.
    Gustafsson, L. et al. Retention forestry to maintain multifunctional forests: a world perspective. Bioscience 62, 633–645 (2012).
    Google Scholar 

    24.
    Schmiegelow, F. K. A., Stepnisky, D. P., Stambaugh, C. A. & Koivula, M. Reconciling salvage logging of boreal forests with a natural-disturbance management model. Conserv. Biol. 20, 971–983 (2006).
    PubMed  Google Scholar 

    25.
    Ministry of Agriculture and Forestry, F. Forest Damages Prevention Act (1087/2013). (2013).

    26.
    De Grandpré, L. et al. Incorporating insect and wind disturbances in a natural disturbance-based management framework for the boreal forest. Forests 9, 1–20 (2018).
    Google Scholar 

    27.
    Chao, A., Colwell, R. K., Gotelli, N. J. & Thorn, S. Proportional mixture of two rarefaction/extrapolation curves to forecast biodiversity changes under landscape transformation. Ecol. Lett. 22, 1913–1922 (2019).
    PubMed  Google Scholar 

    28.
    Anderson, M. J. et al. Navigating the multiple meanings of β diversity: a roadmap for the practicing ecologist. Ecol. Lett. 14, 19–28 (2011).
    ADS  PubMed  Google Scholar 

    29.
    Dornelas, M. et al. Assemblage time series reveal biodiversity change but not systematic loss. Science 344, 296–299 (2014).
    ADS  CAS  PubMed  Google Scholar 

    30.
    Thorn, S. et al. Rare species, functional groups, and evolutionary lineages drive successional trajectories in disturbed forests. Ecology 0, 1–8 (2020).
    Google Scholar 

    31.
    Hyvärinen, E., Kouki, J. & Martikainen, P. Fire and green-tree retention in conservation of red-listed and rare deadwood-dependent beetles in Finnish boreal forests. Conserv. Biol. 20, 1711–1719 (2006).
    PubMed  Google Scholar 

    32.
    Fedrowitz, K. et al. Can retention forestry help conserve biodiversity? A meta-analysis. J. Appl. Ecol. 51, 1669–1679 (2014).
    PubMed  PubMed Central  Google Scholar 

    33.
    Entling, W., Schmidt, M. H., Bacher, S., Brandl, R. & Nentwig, W. Niche properties of Central European spiders: Shading, moisture and the evolution of the habitat niche. Glob. Ecol. Biogeogr. 16, 440–448 (2007).
    Google Scholar 

    34.
    Swanson, M. E. et al. The forgotten stage of forest succession: early-successional ecosystems on forest sites. Front. Ecol. Environ. 9, 117–125 (2011).
    Google Scholar 

    35.
    Lindenmayer, D. B. & Ough, K. Salvage logging in the montane ash eucalypt forests of the Central Highlands of Victoria and its potential impacts on biodiversity. Conserv. Biol. 20, 1005–1015 (2006).
    CAS  PubMed  Google Scholar 

    36.
    Banks-Leite, C. et al. Assessing the utility of statistical adjustments for imperfect detection in tropical conservation science. J. Appl. Ecol. 51, 849–859 (2014).
    PubMed  PubMed Central  Google Scholar 

    37.
    Kortmann, M. et al. Beauty and the beast: how a bat utilizes forests shaped by outbreaks of an insect pest. Anim. Conserv. 21, 21–30 (2018).
    Google Scholar 

    38.
    Mikoláš, M. et al. Mixed-severity natural disturbances promote the occurrence of an endangered umbrella species in primary forests. Ecol. Manag. 405, 210–218 (2017).
    Google Scholar 

    39.
    Leverkus, A. B., Gustafsson, L., Rey Benayas, J. M. & Castro, J. Does post-disturbance salvage logging affect the provision of ecosystem services? A systematic review protocol. Environ. Evid. 4, 16 (2015).
    Google Scholar 

    40.
    Hutto, R. L. & Young, J. Regional landbird monitoring: perspectives from the Northern Rocky Mountains. Wildl. Soc. Bull. 30, 738–750 (2002).
    Google Scholar 

    41.
    Zmihorski, M. The effect of windthrow and its management on breeding bird communities in a managed forest. Biodivers. Conserv. 19, 1871–1882 (2010).
    Google Scholar 

    42.
    Thorn, S. et al. Changes in the dominant assembly mechanism drive species loss caused by declining resources. Ecol. Lett. 19, 163–170 (2016).
    MathSciNet  PubMed  Google Scholar 

    43.
    Leverkus, A. B. et al. Salvage logging effects on regulating and supporting ecosystem services–a systematic map. Can. J. Res. 18, 1–18 (2018).
    Google Scholar 

    44.
    Mehr, M., Brandl, R., Kneib, T. & Müller, J. The effect of bark beetle infestation and salvage logging on bat activity in a national park. Biodivers. Conserv. 21, 2775–2786 (2012).
    Google Scholar 

    45.
    Fontaine, J. B., Donato, D. C., Robinson, W. D., Law, B. E. & Kauffman, J. B. Bird communities following high-severity fire: response to single and repeat fires in a mixed-evergreen forest, Oregon, USA. Ecol. Manag. 257, 1496–1504 (2009).
    Google Scholar 

    46.
    Cahall, R. E. & Hayes, J. P. Influences of postfire salvage logging on forest birds in the Eastern Cascades, Oregon, USA. Ecol. Manag. 257, 1119–1128 (2009).
    Google Scholar 

    47.
    Castro, J., Moreno-Rueda, G. & Hódar, J. Experimental test of postfire management in pine forests: impact of salvage logging versus partial cutting and nonintervention on bird-species assemblages. Conserv. Biol. 24, 810–819 (2010).
    PubMed  Google Scholar 

    48.
    Rost, J., Clavero, M., Brotons, L. & Pons, P. The effect of postfire salvage logging on bird communities in Mediterranean pine forests: the benefits for declining species. J. Appl. Ecol. 49, 644–651 (2012).
    Google Scholar 

    49.
    Zmihorski, M. et al. Early post-fire bird community in European boreal forest: comparing salvage-logged with non-intervention areas. Glob. Ecol. Conserv. 18, e00636 (2019).
    Google Scholar 

    50.
    Choi, C. Y., Lee, E. J., Nam, H. Y. & Lee, W. S. Effects of postfire logging on bird populations and communities in burned forests. J. Korean . Soc. 96, 115–123 (2007).
    Google Scholar 

    51.
    Lee, E.-J., Lee, W.-S., Son, S. H. & Rhim, S.-J. Differences in bird communities in postfire silvicultural practices stands within pine forest of South Korea. Landsc. Ecol. Eng. 7, 137–143 (2011).
    Google Scholar 

    52.
    Koivula, M. & Spence, J. R. Effects of post-fire salvage logging on boreal mixed-wood ground beetle assemblages (Coleoptera, Carabidae). Ecol. Manag. 236, 102–112 (2006).
    Google Scholar 

    53.
    Wermelinger, B. et al. Impact of windthrow and salvage-logging on taxonomic and functional diversity of forest arthropods. Ecol. Manag. 391, 9–18 (2017).
    Google Scholar 

    54.
    Hernández-Hernández, R., Castro, J., Del Arco Aguilar, M., Fernández-López, Á. B. & González-Mancebo, J. M. Post-fire salvage logging imposes a new disturbance that retards succession: the case of bryophyte communities in a Macaronesian laurel forest. Forests 8, 1–16 (2017).
    Google Scholar 

    55.
    Thorn, S. et al. Guild-specific responses of forest Lepidoptera highlight conservation-oriented forest management – implications from conifer-dominated forests. Ecol. Manag. 337, 41–47 (2015).
    Google Scholar 

    56.
    Durska, E. Effects of disturbances on scuttle flies (Diptera: Phoridae) in Pine Forests. Biodivers. Conserv. 22, 1991–2021 (2013).
    Google Scholar 

    57.
    Donato, D. C., Fontaine, J. B., Kauffman, J. B., Robinson, D. & Law, B. E. Fuel mass and forest structure following stand-replacement fire and post-fire logging in a mixed-evergreen forest. Int. J. Wildl. Fire 22, 652–666 (2013).
    Google Scholar 

    58.
    Kurulok, S. E. & Macdonald, S. E. Impacts of postfire salvage logging on understory plant communities of the boreal mixedwood forest 2 and 34 years after disturbance. Can. J. Res. 37, 2637–2651 (2007).
    Google Scholar 

    59.
    Macdonald, S. E. Effects of partial post-fire salvage harvesting on vegetation communities in the boreal mixedwood forest region of northeastern Alberta, Canada. Ecol. Manag. 239, 21–31 (2007).
    Google Scholar 

    60.
    Fornwalt, P. J. et al. Short-term understory plant community responses to salvage logging in beetle-affected lodgepole pine forests. Ecol. Manag. 409, 84–93 (2018).
    Google Scholar 

    61.
    Waldron, K., Ruel, J.-C., Gauthier, S., De Grandpré, L. & Peterson, C. J. Effects of post-windthrow salvage logging on microsites, plant composition and regeneration. Appl. Veg. Sci. 17, 323–337 (2014).
    Google Scholar 

    62.
    Leverkus, A. B., Lorite, J., Navarro, F. B., Sánchez-Cañete, E. P. & Castro, J. Post-fire salvage logging alters species composition and reduces cover, richness, and diversity in Mediterranean plant communities. J. Environ. Manag. 133, 323–331 (2014).
    Google Scholar 

    63.
    Chao, A. et al. Rarefaction and extrapolation with Hill numbers: a framework for sampling and estimation in species diversity studies. Ecol. Monogr. 84, 45–67 (2014).
    Google Scholar 

    64.
    Colwell, R. K. et al. Models and estimators linking individual-based and sample-based rarefaction, extrapolation and comparison of assemblages. J. Plant Ecol. 5, 3–21 (2012).
    Google Scholar 

    65.
    Wood, S. N., Pya, N. & Säfken, B. Smoothing parameter and model selection for general smooth models. J. Am. Stat. Assoc. 111, 1548–1563 (2016).
    MathSciNet  CAS  Google Scholar 

    66.
    Dornelas, M. et al. BioTIME: a database of biodiversity time series for the Anthropocene. Glob. Ecol. Biogeogr. 27, 760–786 (2018).
    PubMed  PubMed Central  Google Scholar 

    67.
    Olson, D. M. et al. Terrestrial ecoregions of the world: a new map of life on earth. Bioscience 51, 933 (2001).
    Google Scholar  More

  • in

    The use of light spectrum blocking films to reduce populations of Drosophila suzukii Matsumura in fruit crops

    1.
    Asplen, M. K. et al. Invasion biology of spotted wing drosophila (Drosophila suzukii): a global perspective and future priorities. J. Pest Sci. 88, 469–494 (2015).
    Google Scholar 
    2.
    Lee, J. C. et al. In focus: spotted wing drosophila, Drosophila suzukii, across perspectives. Pest Manag. Sci. 67, 1349–1351 (2011).
    CAS  PubMed  Google Scholar 

    3.
    Bolda, M. P., Goodhue, R. E. & Zalom, F. G. Spotted wing drosophila: potential economic impact of a newly established pest. Agric. Resour. Econ. Updat. 13, 5–8 (2010).
    Google Scholar 

    4.
    Haviland, D. R. & Beers, E. H. Chemical control programs for Drosophila suzukii that comply with international limitations on pesticide residues for exported sweet cherries. J. Integr. Pest Manag. 3, F1–F6 (2012).
    Google Scholar 

    5.
    Van Timmeren, S., Mota-Sanchez, D., Wise, J. C. & Isaacs, R. Baseline susceptibility of spotted wing drosophila (Drosophila suzukii) to four key insecticide classes. Pest Manag. Sci. 74, 78–87 (2018).
    PubMed  Google Scholar 

    6.
    Gress, B. E. & Zalom, F. G. Identification and risk assessment of spinosad resistance in a California population of Drosophila suzukii. Pest Manag. Sci. 75, 1270–1276 (2019).
    CAS  PubMed  Google Scholar 

    7.
    Bale, J., Van Lenteren, J. & Bigler, F. Biological control and sustainable food production. Philos. Trans. R. Soc. B: Biol. Sci. 363, 761–776 (2008).

    8.
    Iglesias, L. E., Nyoike, T. W. & Liburd, O. E. Effect of trap design, bait type, and age on captures of Drosophila suzukii (Diptera: Drosophilidae) in berry crops. J. Econ. Entomol. 107, 1508–1518 (2014).
    PubMed  Google Scholar 

    9.
    Tonina, L. et al. Comparison of attractants for monitoring Drosophila suzukii in sweet cherry orchards in Italy. J. Appl. Entomol. 142, 18–25 (2018).
    CAS  Google Scholar 

    10.
    Rajapakse, N. C. & Kelly, J. W. Regulation of chrysanthemum growth by spectral filters. J. Am. Soc. for Hortic. Sci. 117, 481–485 (1992).
    Google Scholar 

    11.
    Van Haeringen, C. et al. The development of solid spectral filters for the regulation of plant growth. Photochem. Photobiol. 67, 407–413 (1998).
    Google Scholar 

    12.
    West, J. et al. Spectral filters for the control of Botrytis cinerea. Ann. Appl. Biol. 136, 115–120 (2000).
    ADS  Google Scholar 

    13.
    Antignus, Y., Mor, N., Ben Joseph, R., Lapidot, M. & Cohen, S. Ultraviolet-absorbing plastic sheets protect crops from insect pests and from virus diseases vectored by insects. Environ. Entomol. 25, 919–924 (1996).
    Google Scholar 

    14.
    Fennell, J. T., Fountain, M. T. & Paul, N. D. Direct effects of protective cladding material on insect pests in crops. Crop. Prot. (2019).

    15.
    Doukas, D. & Payne, C. The use of ultraviolet-blocking films in insect pest management in the UK; effects on naturally occurring arthropod pest and natural enemy populations in a protected cucumber crop. Ann. Appl. Biol. 151, 221–231 (2007).
    Google Scholar 

    16.
    Solaiman, A. H. M., Nishizawa, T., Arefin, S. A., Sarkar, M. D. & Shahjahan, M. Effect of partially UV-blocking films on the growth, yield, pigmentation, and insect control of red amaranth (Amaranthus tricolor). Curr. J. Appl. Sci. Technol. 1–11 (2016).

    17.
    Hardie, R. C. Functional organization of the fly retina. In Progress in Sensory Physiology, 1–79 (Springer, New York, 1985).

    18.
    Schnaitmann, C., Pagni, M. & Reiff, D. F. Color vision in insects: insights from drosophila. J. Comp. Physiol. A 206, 1–16 (2020).
    Google Scholar 

    19.
    Schnaitmann, C., Garbers, C., Wachtler, T. & Tanimoto, H. Color discrimination with broadband photoreceptors. Curr. Biol. 23, 2375–2382 (2013).
    CAS  PubMed  Google Scholar 

    20.
    Wardill, T. J. et al. Multiple spectral inputs improve motion discrimination in the drosophila visual system. Science 336, 925–931 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    21.
    Schümperli, R. A. Evidence for colour vision in Drosophila melanogaster through spontaneous phototactic choice behaviour. J. Comp. Physiol. A 86, 77–94 (1973).
    Google Scholar 

    22.
    Bernard, G. D. & Stavenga, D. G. Spectral sensitivities of retinular cells measured in intact, living flies by an optical method. J. Comp. Physiol. 134, 95–107 (1979).
    Google Scholar 

    23.
    Hardie, R. C. Polarization vision: Drosophila enters the arena. Curr. Biol. 22, R12–R14 (2012).
    CAS  PubMed  Google Scholar 

    24.
    Zhu, E. Y., Guntur, A. R., He, R., Stern, U. & Yang, C.-H. Egg-laying demand induces aversion of UV light in drosophila females. Curr. Biol. 24, 2797–2804 (2014).
    CAS  PubMed  PubMed Central  Google Scholar 

    25.
    Kane, E. A. et al. Sensorimotor structure of drosophila larva phototaxis. Proc. Natl. Acad. Sci. 110, E3868–E3877 (2013).
    ADS  CAS  PubMed  Google Scholar 

    26.
    Kelber, A. & Henze, M. J. Colour vision: parallel pathways intersect in drosophila. Curr. Biol. 23, R1043–R1045 (2013).
    CAS  PubMed  Google Scholar 

    27.
    Rice, K. B., Short, B. D., Jones, S. K. & Leskey, T. C. Behavioral responses of Drosophila suzukii (Diptera: Drosophilidae) to visual stimuli under laboratory, semifield, and field conditions. Environ. Entomol. 45, 1480–1488 (2016).
    PubMed  Google Scholar 

    28.
    Kirkpatrick, D., McGhee, P., Hermann, S., Gut, L. & Miller, J. Alightment of spotted wing drosophila (Diptera: Drosophilidae) on odorless disks varying in color. Environ. Entomol. 45, 185–191 (2016).
    CAS  PubMed  Google Scholar 

    29.
    Little, C. M., Chapman, T. W. & Hillier, N. K. Effect of color and contrast of highbush blueberries to host-finding behavior by drosophila suzukii (Diptera: Drosophilidae). Environ. Entomol. 47, 1242–1251 (2018).
    CAS  PubMed  Google Scholar 

    30.
    Little, C. M., Rizzato, A. R., Charbonneau, L., Chapman, T. & Hillier, N. K. Color preference of the spotted wing drosophila. Drosophila suzukii. Sci. Rep. 9, 1–12 (2019).
    Google Scholar 

    31.
    Yamaguchi, S., Wolf, R., Desplan, C. & Heisenberg, M. Motion vision is independent of color in drosophila. Proc. Natl. Acad. Sci. 105, 4910–4915 (2008).
    ADS  CAS  PubMed  Google Scholar 

    32.
    Paulk, A., Millard, S. S. & van Swinderen, B. Vision in drosophila: seeing the world through a model’s eyes. Annu. Rev. Entomol. 58, 313–332 (2013).
    CAS  PubMed  Google Scholar 

    33.
    Humberg, T.-H. & Sprecher, S. G. Age-and wavelength-dependency of drosophila larval phototaxis and behavioral responses to natural lighting conditions. Front. Behav. Neurosci. 11, 66 (2017).
    PubMed  PubMed Central  Google Scholar 

    34.
    Cronin, T. W. & Bok, M. J. Photoreception and vision in the ultraviolet. J. Exp. Biol. 219, 2790–2801 (2016).
    PubMed  Google Scholar 

    35.
    Stone, T., Mangan, M., Ardin, P., Webb, B. et al. Sky segmentation with ultraviolet images can be used for navigation. In Robotics: Science and Systems (2014).

    36.
    Kirk, R., Cielniak, G. & Mangan, M. L* a* b* fruits: A rapid and robust outdoor fruit detection system combining bio-inspired features with one-stage deep learning networks. Sensors 20, 275 (2020).
    Google Scholar 

    37.
    Clymans, R. et al. Olfactory preference of Drosophila suzukii shifts between fruit and fermentation cues over the season: effects of physiological status. Insects 10, 200 (2019).
    PubMed Central  Google Scholar 

    38.
    Keesey, I. W., Knaden, M. & Hansson, B. S. Olfactory specialization in Drosophila suzukii supports an ecological shift in host preference from rotten to fresh fruit. J. Chem. Ecol. 41, 121–128 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    39.
    Kumar, P. & Poehling, H.-M. UV-blocking plastic films and nets influence vectors and virus transmission on greenhouse tomatoes in the humid tropics. Environ. Entomol. 35, 1069–1082 (2006).
    Google Scholar 

    40.
    Legarrea, S., Karnieli, A., Fereres, A. & Weintraub, P. G. Comparison of UV-absorbing nets in pepper crops: Spectral properties, effects on plants and pest control. Photochem. Photobiol. 86, 324–330 (2010).
    CAS  PubMed  Google Scholar 

    41.
    Costa, H. S. & Robb, K. L. Effects of ultraviolet-absorbing greenhouse plastic films on flight behavior of Bemisia argentifolii (homoptera: Aleyrodidae) and Frankliniella occidentalis (Thysanoptera: Thripidae). J. Econ. Entomol. 92, 557–562 (1999).
    Google Scholar 

    42.
    Chyzik, R., Dobrinin, S. & Antignus, Y. Effect of a UV-deficient environment on the biology and flight activity of Myzus persicae and its hymenopterous parasite Aphidius matricariae. Phytoparasitica 31, 467–477 (2003).
    Google Scholar 

    43.
    Costa, H., Robb, K. & Wilen, C. Field trials measuring the effects of ultraviolet-absorbing greenhouse plastic films on insect populations. J. Econ. Entomol. 95, 113–120 (2002).
    CAS  PubMed  Google Scholar 

    44.
    Dáder, B., Gwynn-Jones, D., Moreno, A., Winters, A. & Fereres, A. Impact of uv-a radiation on the performance of aphids and whiteflies and on the leaf chemistry of their host plants. J. Photochem. Photobiol. B: Biol. 138, 307–316 (2014).
    Google Scholar 

    45.
    Díaz, B. M., Biurrún, R., Moreno, A., Nebreda, M. & Fereres, A. Impact of ultraviolet-blocking plastic films on insect vectors of virus diseases infesting crisp lettuce. HortScience 41, 711–716 (2006).
    Google Scholar 

    46.
    Kuhlmann, F. & Müller, C. Development-dependent effects of UV radiation exposure on broccoli plants and interactions with herbivorous insects. Environ. Exp. Bot. 66, 61–68 (2009).
    CAS  Google Scholar 

    47.
    Paul, N. D. et al. Ecological responses to UV radiation: interactions between the biological effects of UV on plants and on associated organisms. Physiol. Plantarum 145, 565–581 (2012).
    CAS  Google Scholar 

    48.
    Sal, J. et al. Influence of UV-absorbing nets in the population of Macrosiphum euphorbiae Thomas (Homoptera: Aphididae) and the parasitoid Aphidius ervi (Haliday) (Hymenoptera: Aphidiidae) in lettuce crops. In Proceedings of Third International Symposium Biological Control Arthropods, Christ Church, New Zealand, 329–337 (2009).

    49.
    Legarrea, S., Weintraub, P., Plaza, M., Viñuela, E. & Fereres, A. Dispersal of aphids, whiteflies and their natural enemies under photoselective nets. Biocontrol 57, 523–532 (2012).
    Google Scholar 

    50.
    Legarrea, S. et al. Dynamics of nonpersistent aphid-borne viruses in lettuce crops covered with UV-absorbing nets. Virus Res. 165, 1–8 (2012).
    CAS  PubMed  Google Scholar 

    51.
    Legarrea, S. et al. Diminished uv radiation reduces the spread and population density of Macrosiphum euphorbiae (Thomas) [Hemiptera: Aphididae] in lettuce crops. Hortic. Sci. 39, 74–80 (2012).
    Google Scholar 

    52.
    Dáder, B., Moreno, A., Gwynn-Jones, D., Winters, A. & Fereres, A. Aphid orientation and performance in glasshouses under different UV-a/UV-b radiation regimes. Entomol. Exp. et Appl. 163, 344–353 (2017).
    Google Scholar 

    53.
    El-Aal, H. A. A., Rizk, A. M. & Mousa, I. E. Evaluation of new greenhouse covers with modified light regime to control cotton aphid and cucumber (Cucumis sativus L.) productivity. Crop. Prot. 107, 64–70 (2018).
    Google Scholar 

    54.
    Kigathi, R. & Poehling, H.-M. UV-absorbing films and nets affect the dispersal of western flower thrips, Frankliniella occidentalis (Thysanoptera: Thripidae). J. Appl. Entomol. 136, 761–771 (2012).
    Google Scholar 

    55.
    Bueno, E. et al. Response of wild spotted wing drosophila (Drosophila suzukii) to microbial volatiles. J. Chem. Ecol. 1–11 (2019).

    56.
    Renkema, J. M., Buitenhuis, R. & Hallett, R. H. Reduced Drosophila suzukii infestation in berries using deterrent compounds and laminate polymer flakes. Insects 8, 117 (2017).
    PubMed Central  Google Scholar 

    57.
    Erland, L. A., Rheault, M. R. & Mahmoud, S. S. Insecticidal and oviposition deterrent effects of essential oils and their constituents against the invasive pest Drosophila suzukii (Matsumura) (Diptera: Drosophilidae). Crop. Prot. 78, 20–26 (2015).
    CAS  Google Scholar 

    58.
    Wallingford, A. K., Cha, D. H. & Loeb, G. M. Evaluating a push–pull strategy for management of Drosophila suzukii Matsumura in red raspberry. Pest Manag. Sci. 74, 120–125 (2018).
    CAS  PubMed  Google Scholar 

    59.
    Smirle, M. J., Zurowski, C. L., Ayyanath, M.-M., Scott, I. M. & MacKenzie, K. E. Laboratory studies of insecticide efficacy and resistance in Drosophila suzukii (Matsumura) (Diptera: Drosophilidae) populations from British Columbia, Canada. Pest Manag. Sci. 73, 130–137 (2017).
    CAS  PubMed  Google Scholar 

    60.
    Shaw, B., Brain, P., Wijnen, H. & Fountain, M. T. Implications of sub-lethal rates of insecticides and daily time of application on Drosophila suzukii lifecycle. Crop. Prot. 121, 182–194 (2019).
    CAS  Google Scholar 

    61.
    van der Blom, J. Applied entomology in spanish greenhouse horticulture. Proc. Neth. Entomol. Soc. Meet 21, 9–17 (2010).
    Google Scholar 

    62.
    Fingerman, M. & Brown, F. A. A “purkinje shift” in insect vision. Science 116, 171–172 (1952).
    ADS  CAS  PubMed  Google Scholar 

    63.
    Meier, U. Phenological growth stages. In Phenology: An Integrative Environmental Science, 269–283 (Springer, New York, 2003).

    64.
    Doukas, D. & Payne, C. C. Greenhouse whitefly (Homoptera: Aleyrodidae) dispersal under different UV-light environments. J. Econ. Entomol. 100, 389–397 (2014).
    Google Scholar 

    65.
    Palanca, L., Gaskett, A. C., Günther, C. S., Newcomb, R. D. & Goddard, M. R. Quantifying variation in the ability of yeasts to attract Drosophila melanogaster. PLoS ONE 8 (2013). More

  • in

    Root pathogen diversity and composition varies with climate in undisturbed grasslands, but less so in anthropogenically disturbed grasslands

    1.
    Mordecai EA. Pathogen impacts on plant communities: unifying theory, concepts, and empirical work. Ecol Monogr. 2011;81:429–41.
    Article  Google Scholar 
    2.
    Bever JD, Mangan SA, Alexander HM. Maintenance of plant species diversity by pathogens. Annu Rev Ecol Evol Syst. 2015;46:305–25.
    Article  Google Scholar 

    3.
    van der Heijden MG, Bardgett RD, van Straalen NM. The unseen majority: soil microbes as drivers of plant diversity and procductivity in terrestrial ecosystems. Ecol Lett. 2008;11:296–310.
    PubMed  Article  PubMed Central  Google Scholar 

    4.
    Mangan SA, Schnitzer SA, Herre EA, Mack KM, Valencia MC, Sanchez EI, et al. Negative plant-soil feedback predicts tree-species relative abundance in a tropical forest. Nature. 2010;466:752–5.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Comita LS, Muller-Landau HC, Aguilar S, Hubbell SP. Asymmetric density dependence shapes species abundances in a tropical tree community. Science. 2010;329:330–2.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    6.
    Janzen DH. Herbivores and the number of tree species in tropical forests. Am Naturalist. 1970;104:501–28.
    Article  Google Scholar 

    7.
    Connell JH. On the role of natural enemies in preventing competitive exclusion in some marine animals and in rain forest trees. Dyn Popul. 1971;298:312.
    Google Scholar 

    8.
    Augspurger CK. Seedling survival of tropical tree species: interactions of dispersal distance, light‐gaps, and pathogens. Ecology. 1984;65:1705–12.
    Article  Google Scholar 

    9.
    Van der Putten WH, Bardgett RD, Bever JD, Bezemer TM, Casper BB, Fukami T, et al. Plant–soil feedbacks: the past, the present and future challenges. J Ecol. 2013;101:265–76.
    Article  Google Scholar 

    10.
    Eppinga MB, Baudena M, Johnson DJ, Jiang J, Mack KM, Strand AE, et al. Frequency-dependent feedback constrains plant community coexistence. Nat Ecol evolution. 2018;2:1403.
    Article  Google Scholar 

    11.
    Crawford KM, Bauer JT, Comita LS, Eppinga MB, Johnson DJ, Mangan SA, et al. When and where plant‐soil feedback may promote plant coexistence: a meta‐analysis. Ecol Lett. 2019;22:1274–84.
    PubMed  Google Scholar 

    12.
    Callaway RM, Thelen GC, Rodriguez A, Holben WE. Soil biota and exotic plant invasion. Nature. 2004;427:731–3.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    13.
    Mitchell CE, Agrawal AA, Bever JD, Gilbert GS, Hufbauer RA, Klironomos JN, et al. Biotic interactions and plant invasions. Ecol Lett. 2006;9:726–40.
    PubMed  Article  PubMed Central  Google Scholar 

    14.
    Fierer N, Strickland MS, Liptzin D, Bradford MA, Cleveland CC. Global patterns in belowground communities. Ecol Lett. 2009;12:1238–49.
    PubMed  Article  PubMed Central  Google Scholar 

    15.
    Rousk J, Bååth E, Brookes PC, Lauber CL, Lozupone C, Caporaso JG, et al. Soil bacterial and fungal communities across a pH gradient in an arable soil. ISME J. 2010;4:1340.
    PubMed  Article  PubMed Central  Google Scholar 

    16.
    Thomson BC, Tisserant E, Plassart P, Uroz S, Griffiths RI, Hannula SE, et al. Soil conditions and land use intensification effects on soil microbial communities across a range of European field sites. Soil Biol Biochem. 2015;88:403–13.
    CAS  Article  Google Scholar 

    17.
    Van Agtmaal M, Straathof A, Termorshuizen A, Teurlincx S, Hundscheid M, Ruyters S, et al. Exploring the reservoir of potential fungal plant pathogens in agricultural soil. Appl Soil Ecol. 2017;121:152–60.
    Article  Google Scholar 

    18.
    Rincón A, Santamaría‐Pérez B, Rabasa SG, Coince A, Marçais B, Buée M. Compartmentalized and contrasted response of ectomycorrhizal and soil fungal communities of Scots pine forests along elevation gradients in France and Spain. Environ Microbiol. 2015;17:3009–24.
    PubMed  Article  PubMed Central  Google Scholar 

    19.
    Newsham KK, Hopkins DW, Carvalhais LC, Fretwell PT, Rushton SP, O’Donnell AG, et al. Relationship between soil fungal diversity and temperature in the maritime Antarctic. Nat Clim Change. 2016;6:182.
    Article  Google Scholar 

    20.
    Zhou J, Deng Y, Shen L, Wen C, Yan Q, Ning D, et al. Temperature mediates continental-scale diversity of microbes in forest soils. Nat Commun. 2016;7:12083.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    21.
    Tedersoo L, Bahram M, Põlme S, Kõljalg U, Yorou NS, Wijesundera R, et al. Global diversity and geography of soil fungi. Science. 2014;346:1256688.
    PubMed  Article  CAS  Google Scholar 

    22.
    McGuire KL, Fierer N, Bateman C, Treseder KK, Turner BL. Fungal community composition in neotropical rain forests: the influence of tree diversity and precipitation. Micro Ecol. 2012;63:804–12.
    Article  Google Scholar 

    23.
    Oliverio AM, Geisen S, Delgado-Baquerizo M, Maestre FT, Turner BL, Fierer N. The global-scale distributions of soil protists and their contributions to belowground systems. Sci Adv. 2020;6:eaax8787.
    PubMed  PubMed Central  Article  Google Scholar 

    24.
    Talley SM, Coley PD, Kursar TA. The effects of weather on fungal abundance and richness among 25 communities in the Intermountain West. BMC Ecol. 2002;2:7.
    PubMed  PubMed Central  Article  Google Scholar 

    25.
    Spear ER. Phylogenetic relationships and spatial distributions of putative fungal pathogens of seedlings across a rainfall gradient in Panama. Fungal Ecol. 2017;26:65–73.
    Article  Google Scholar 

    26.
    Geml J, Pastor N, Fernandez L, Pacheco S, Semenova TA, Becerra AG, et al. Large‐scale fungal diversity assessment in the Andean Yungas forests reveals strong community turnover among forest types along an altitudinal gradient. Mol Ecol. 2014;23:2452–72.
    CAS  PubMed  Article  Google Scholar 

    27.
    Rojas JA, Jacobs JL, Napieralski S, Karaj B, Bradley CA, Chase T, et al. Oomycete species associated with soybean seedlings in North America—Part II: diversity and ecology in relation to environmental and edaphic factors. Phytopathology. 2017;107:293–304.
    PubMed  Article  Google Scholar 

    28.
    van West P, Appiah AA, Gow NA. Advances in research on oomycete root pathogens. Physiol Mol plant Pathol. 2003;62:99–113.
    Article  Google Scholar 

    29.
    Oehl F, Sieverding E, Ineichen K, Mäder P, Boller T, Wiemken A. Impact of land use intensity on the species diversity of arbuscular mycorrhizal fungi in agroecosystems of Central Europe. Appl Environ Microbiol. 2003;69:2816–24.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    30.
    House GL, Bever JD. Disturbance reduces the differentiation of mycorrhizal fungal communities in grasslands along a precipitation gradient. Ecol Appl. 2018;28:736–48.
    PubMed  Article  PubMed Central  Google Scholar 

    31.
    Trenberth KE, Dai A, Van Der Schrier G, Jones PD, Barichivich J, Briffa KR, et al. Global warming and changes in drought. Nat Clim Change. 2014;4:17.
    Article  Google Scholar 

    32.
    IPCC. Climate change 2014: synthesis report. Switzerland: IPCC Geneva; 2014. p. 151.
    Google Scholar 

    33.
    Zhang N, Wan S, Guo J, Han G, Gutknecht J, Schmid B, et al. Precipitation modifies the effects of warming and nitrogen addition on soil microbial communities in northern Chinese grasslands. Soil Biol Biochem. 2015;89:12–23.
    CAS  Article  Google Scholar 

    34.
    Sheik CS, Beasley WH, Elshahed MS, Zhou X, Luo Y, Krumholz LR. Effect of warming and drought on grassland microbial communities. ISME J. 2011;5:1692.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    35.
    Wu Z, Dijkstra P, Koch GW, Peñuelas J, Hungate BA. Responses of terrestrial ecosystems to temperature and precipitation change: a meta‐analysis of experimental manipulation. Glob Change Biol. 2011;17:927–42.
    Article  Google Scholar 

    36.
    Ihrmark K, Bödeker IT, Cruz-Martinez K, Friberg H, Kubartova A, Schenck J, et al. New primers to amplify the fungal ITS2 region–evaluation by 454-sequencing of artificial and natural communities. FEMS Microbiol Ecol. 2012;82:666–77.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    37.
    White TJ, Bruns T, Lee S, Taylor J. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. PCR Protoc Guide Methods Appl. 1990;18:315–22.
    Google Scholar 

    38.
    Schoch CL, Seifert KA, Huhndorf S, Robert V, Spouge JL, Levesque CA, et al. Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for fungi. Proc Natl Acad Sci. 2012;109:6241–6.

    39.
    Oliver AK, Callaham MA Jr, Jumpponen A. Soil fungal communities respond compositionally to recurring frequent prescribed burning in a managed southeastern US forest ecosystem. For Ecol Manag. 2015;345:1–9.
    Article  Google Scholar 

    40.
    Riit T, Tedersoo L, Drenkhan R, Runno-Paurson E, Kokko H, Anslan S. Oomycete-specific ITS primers for identification and metabarcoding. MycoKeys. 2016;14:17.
    Article  Google Scholar 

    41.
    Caporaso JG, Kuczynski J, Stombaugh J, Bittinger K, Bushman FD, Costello EK, et al. QIIME allows analysis of high-throughput community sequencing data. Nat Methods. 2010;7:335.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    42.
    Kõljalg U, Nilsson RH, Abarenkov K, Tedersoo L, Taylor AF, Bahram M, et al. Towards a unified paradigm for sequence‐based identification of fungi. Mol Ecol. 2013;22:5271–7.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    43.
    Lindahl BD, Nilsson RH, Tedersoo L, Abarenkov K, Carlsen T, Kjøller R, et al. Fungal community analysis by high‐throughput sequencing of amplified markers–a user’s guide. N. Phytologist. 2013;199:288–99.
    CAS  Article  Google Scholar 

    44.
    Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15:550.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    45.
    Wang Q, Garrity GM, Tiedje JM, Cole JR. Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl Environ Microbiol. 2007;73:5261–7.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    46.
    James TY, Kauff F, Schoch CL, Matheny PB, Hofstetter V, Cox CJ, et al. Reconstructing the early evolution of fungi using a six-gene phylogeny. Nature 2006;443:818.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    47.
    Nguyen NH, Song Z, Bates ST, Branco S, Tedersoo L, Menke J, et al. FUNGuild: an open annotation tool for parsing fungal community datasets by ecological guild. Fungal Ecol. 2016;20:241–8.
    Article  Google Scholar 

    48.
    Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W, et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997;25:3389–402.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    49.
    Rujirawat T, Patumcharoenpol P, Lohnoo T, Yingyong W, Kumsang Y, Payattikul P, et al. Probing the phylogenomics and putative pathogenicity genes of pythium insidiosum by oomycete genome analyses. Sci Rep. 2018;8:4135.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    50.
    Team RC. R: A language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing; 2016. p. 2017.
    Google Scholar 

    51.
    Helmus MR, Bland TJ, Williams CK, Ives AR. Phylogenetic measures of biodiversity. Am Naturalist. 2007;169:E68–83.
    Article  Google Scholar 

    52.
    Stamatakis A. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 2006;22:2688–90.
    CAS  PubMed  Article  Google Scholar 

    53.
    Nilsson RH, Kristiansson E, Ryberg M, Hallenberg N, Larsson K-H. Intraspecific ITS variability in the kingdom Fungi as expressed in the international sequence databases and its implications for molecular species identification. Evolut Bioinforma. 2008;4:EBO. S653.
    Article  Google Scholar 

    54.
    Pearse WD, Cadotte MW, Cavender-Bares J, Ives AR, Tucker CM, Walker SC, et al. Pez: Phylogenetics for the environmental sciences. Bioinformatics. 2015;31:2888–90.
    CAS  PubMed  Article  Google Scholar 

    55.
    Fitzpatrick DA, Logue ME, Stajich JE, Butler G. A fungal phylogeny based on 42 complete genomes derived from supertree and combined gene analysis. BMC Evolut Biol. 2006;6:99.
    Article  CAS  Google Scholar 

    56.
    Lauber CL, Strickland MS, Bradford MA, Fierer N. The influence of soil properties on the structure of bacterial and fungal communities across land-use types. Soil Biol Biochem. 2008;40:2407–15.
    CAS  Article  Google Scholar 

    57.
    Chaudhary VB, O’Dell TE, Rillig MC, Johnson NC. Multiscale patterns of arbuscular mycorrhizal fungal abundance and diversity in semiarid shrublands. Fungal Ecol. 2014;12:32–43.
    Article  Google Scholar 

    58.
    Morisita M. Measuring of interspecific association and similarity between communities. Mem Fac Sci Kyushu Univ Ser E. 1959;3:65–80.
    Google Scholar 

    59.
    Oksanen J, Blanchet FG, Kindt R, Legendre P, Minchin PR, O’hara R, et al. Package ‘vegan’. Community ecology package, version. 2013;2.

    60.
    McMurdie PJ, Holmes S. Waste not, want not: why rarefying microbiome data is inadmissible. PLoS Comput Biol. 2014;10:e1003531.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    61.
    Ochoa‐Hueso R, Collins SL, Delgado‐Baquerizo M, Hamonts K, Pockman WT, Sinsabaugh RL, et al. Drought consistently alters the composition of soil fungal and bacterial communities in grasslands from two continents. Glob Change Biol. 2018;24:2818–27.
    Article  Google Scholar 

    62.
    Dequiedt S, Saby N, Lelievre M, Jolivet C, Thioulouse J, Toutain B, et al. Biogeographical patterns of soil molecular microbial biomass as influenced by soil characteristics and management. Glob Ecol Biogeogr. 2011;20:641–52.
    Article  Google Scholar 

    63.
    Samson FB, Knopf FL, Ostlie WR. Great Plains ecosystems: past, present, and future. Wildl Soc Bull. 2004;32:6–15.
    Article  Google Scholar 

    64.
    Gilbert GS, Webb CO. Phylogenetic signal in plant pathogen-host range. Proc Natl Acad Sci USA. 2007;104:4979–83.
    CAS  PubMed  Article  Google Scholar 

    65.
    Robideau GP, de Cock AW, Coffey MD, Voglmayr H, Brouwer H, Bala K, et al. DNA barcoding of oomycetes with cytochrome c oxidase subunit I and internal transcribed spacer. Mol Ecol Resour. 2011;11:1002–11.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    66.
    Martin LM, Moloney KA, Wilsey BJ. An assessment of grassland restoration success using species diversity components. J Appl Ecol. 2005;42:327–36.
    Article  Google Scholar 

    67.
    Leach MK, Givnish TJ. Ecological determinants of species loss in remnant prairies. Science. 1996;273:1555–8.
    CAS  Article  Google Scholar  More

  • in

    Rapid functional traits turnover in boreal dragonfly communities (Odonata)

    1.
    Bjørnstad, O. N. & Grenfell, B. T. Noisy clockwork: Time series analysis of population fluctuations in animals. Science 293, 638–643 (2001).
    PubMed  Google Scholar 
    2.
    Ricklefs, R. E. & Relyea, R. Ecology, The Economy of Nature 7th edn. (W.H. Freeman & Co., Ltd., New York, 2014).
    Google Scholar 

    3.
    Beisner, B. E., Haydon, D. T. & Cuddington, K. Alternative stable states in ecology. Front. Ecol. Environ. 1, 376–382 (2003).
    Google Scholar 

    4.
    Calijuri, M. D. C., Dos Santos, A. C. A. & Jati, S. Temporal changes in the phytoplankton community structure in a tropical and eutrophic reservoir (Barra Bonita, SP—Brazil). J. Plankton Res. 24, 617–634 (2002).
    CAS  Google Scholar 

    5.
    Allan, E. et al. More diverse plant communities have higher functioning over time due to turnover in complementary dominant species. Proc. Natl. Acad. Sci. U.S.A. 108, 17034–17039 (2011).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    6.
    Magurran, A. E. & Henderson, P. A. Temporal turnover and the maintenance of diversity in ecological assemblages. Philos. Trans. R. Soc. Lond. Ser. B. 365, 3611–3620 (2010).
    Google Scholar 

    7.
    Poff, N. L. et al. Functional trait niches of North American lotic insects: traits-based ecological applications in light of phylogenetic relationships. J. N. Am. Benthol. Soc. 25, 730–755 (2006).
    Google Scholar 

    8.
    Villéger, S., Mason, N. W. & Mouillot, D. New multidimensional functional diversity indices for a multifaceted framework in functional ecology. Ecology 89, 2290–2301 (2008).
    PubMed  Google Scholar 

    9.
    Powney, G. D., Cham, S. S., Smallshire, D. & Isaac, N. J. Trait correlates of distribution trends in the Odonata of Britain and Ireland. PeerJ 3, e1410. https://doi.org/10.7717/peerj.1410 (2015).
    Article  PubMed  PubMed Central  Google Scholar 

    10.
    Moldan, F., Cosby, B. J. & Wright, R. F. Modeling past and future acidification of Swedish lakes. Ambio 42, 577–586 (2013).
    PubMed  PubMed Central  Google Scholar 

    11.
    Levers, C. et al. Drivers of forest harvesting intensity patterns in Europe. For. Ecol. Manag. 315, 160–172 (2014).
    Google Scholar 

    12.
    Cousins, S. A., Auffret, A. G., Lindgren, J. & Tränk, L. Regional-scale land-cover change during the 20th century and its consequences for biodiversity. Ambio 44, 17–27 (2015).
    PubMed Central  Google Scholar 

    13.
    HELCOM, Helsinki Commission. Climate change in the Baltic Sea Area. HELCOM thematic assessment in 2013; https://www.helcom.fi/lists/publications/bsep137.pdf (2013).

    14.
    Daniel, J., Gleason, J. E., Cottenie, K. & Rooney, R. C. Stochastic and deterministic processes drive wetland community assembly across a gradient of environmental filtering. Oikos 128, 1158–1169 (2019).
    Google Scholar 

    15.
    Hassall, C. Odonata as candidate macroecological barometers for global climate change. Fresh Sci. 34, 1040–1049 (2015).
    Google Scholar 

    16.
    Sahlén, G. & Ekestubbe, K. Identification of dragonflies (Odonata) as indicators of general species richness in boreal forest lakes. Biodiv. Cons. 10, 673–690 (2001).
    Google Scholar 

    17.
    Monteiro-Júnior, C. D. S., Juen, L. & Hamada, N. Analysis of urban impacts on aquatic habitats in the central Amazon basin: adult odonates as bioindicators of environmental quality. Ecol. Ind. 48, 303–311 (2015).
    Google Scholar 

    18.
    Suhling, F. et al. Order Odonata. In Ecology and General Biology: Thorp and Covich’s Freshwater Invertebrates (eds Thorp, J. & Rogers, D. C.) 893–932 (Academic Press, New York, 2015).
    Google Scholar 

    19.
    Appelberg, M., Henrikson, B. I., Henrikson, L. & Svedäng, M. Biotic interactions within the littoral community of Swedish forest lakes during acidification. Ambio 22, 290–297 (1993).
    Google Scholar 

    20.
    Al Jawaheri, R. & Sahlén, G. Negative impact of lake liming programmes on the species richness of dragonflies (Odonata): A study from southern Sweden. Hydrobiologia 788, 99–113 (2017).
    Google Scholar 

    21.
    Sahlén, G. Specialists vs generalists in the Odonata–the importance of forest environments in the formation of diverse species pools. In Forests and dragonflies (ed. Cordero Rivera, A.) 153–179 (Pensoft, Sofia, 2006).
    Google Scholar 

    22.
    Dalzochio, M. S., Périco, E., Renner, S. & Sahlén, G. Effect of tree plantations on the functional composition of Odonata species in the highlands of southern Brazil. Hydrobiologia 808, 283–300 (2018).
    Google Scholar 

    23.
    Renner, S., Périco, E., Dalzochio, M. S. & Sahlén, G. Water body type and land cover shape the dragonfly communities (Odonata) in the Pampa biome, Rio Grande do Sul Brazil. J. Insect Cons. 22, 113–125 (2018).
    Google Scholar 

    24.
    Flenner, I. & Sahlén, G. Dragonfly community re-organisation in boreal forest lakes: rapid species turnover driven by climate change?. Insect Conserv. Diver. 1, 169–179 (2008).
    Google Scholar 

    25.
    Ball-Damerow, J. E., M’Gonigle, L. K. & Resh, V. H. Changes in occurrence, richness, and biological traits of dragonflies and damselflies (Odonata) in California and Nevada over the past century. Biodiv. Cons. 23, 2107–2126 (2014).
    Google Scholar 

    26.
    Buisson, L., Grenouillet, G., Villéger, S., Canal, J. & Laffaille, P. Toward a loss of functional diversity in stream fish assemblages under climate change. Global Change Biol. 19, 387–400 (2013).
    ADS  Google Scholar 

    27.
    Angert, A. L. et al. Do species’ traits predict recent shifts at expanding range edges?. Ecol. Lett. 14, 677–689 (2011).
    PubMed  Google Scholar 

    28.
    Lawson, C., Vindenes, Y., Baley, L. & van de Pol, M. Environmental variation and population responses to global change. Ecol. Lett. 18, 724–736 (2015).
    PubMed  Google Scholar 

    29.
    Shimadzu, H., Dornelas, M. & Magurran, A. E. Measuring temporal turnover in ecological communities. Methods Ecol. Evol. 6, 1384–1394 (2015).
    Google Scholar 

    30.
    Jonsson, M. et al. Climate change modifies the size structure of assemblages of emerging aquatic insects. Freshw. Biol. 60, 78–88 (2015).
    Google Scholar 

    31.
    Koch, K., Wagner, C. & Sahlén, G. Farmland versus forest: comparing changes in Odonata species composition in western and eastern Sweden. Insect Cons. Divers. 7, 22–31 (2014).
    Google Scholar 

    32.
    Norling, U. & Sahlén, G. Odonata, dragonflies and damselflies in Aquatic insects of North Europe: a taxonomic handbook, Vol. 2 (ed. Nilsson, A.) 13–65 (Apollo books, 1997).

    33.
    Corbet, P. S. Dragonflies–behaviour and ecology of Odonata (Cornell University Press, Cornell, 1999).
    Google Scholar 

    34.
    Pereira, D. F. G., de Oliveira Junior, J. M. B. & Juen, L. Environmental changes promote larger species of Odonata (Insecta) in Amazonian streams. Ecol. Ind. 98, 179–192 (2019).
    Google Scholar 

    35.
    Johansson, F., Śniegula, S. & Brodin, T. Emergence patterns and latitudinal adaptations in development time of Odonata in north Sweden and Poland. Odonatologica 39, 97–106 (2010).
    Google Scholar 

    36.
    Suhling, I. & Suhling, F. Thermal adaptation affects interactions between a range-expanding and a native odonate species. Freshw. Biol. 58, 705–714 (2013).
    Google Scholar 

    37.
    Atkinson, D. Temperature and organism size: a biological law for ectotherms?. Adv. Ecol. Res. 25, 1–58 (1994).
    Google Scholar 

    38.
    Menéndez, R. How are insects responding to global warming?. Tijdschr. Entomol. 150, 355–364 (2007).
    Google Scholar 

    39.
    Hogue, J. N. & Hawkins, C. P. Morphological variation in adult aquatic insects: Associations with developmental temperature and seasonal growth patterns. J. N. Am. Benth. Soc. 10, 309–321 (1991).
    Google Scholar 

    40.
    Waringer, J. A. & Humpesch, U. H. Embryonic development, larval growth and life cycle of Coenagrion puella (Odonata: Zygoptera) from an Austrian pond. Freshw. Biol. 14, 385–399 (1984).
    Google Scholar 

    41.
    Martens, A. Annual development of Libellula quadrimaculata L in a newly setup pond (Anisoptera: Libellulidae). Notul. Odonatol. 2, 133–134 (1986).
    Google Scholar 

    42.
    Norling, U. Life history patterns in the northern expansion of dragonflies. Adv. Odonatol. 2, 127–156 (1984).
    Google Scholar 

    43.
    Hassall, C., Thompson, D. J., French, G. C. & Harvey, I. F. Historical changes in the phenology of British Odonata are related to climate. Glob. Change Biol. 13, 933–941 (2007).
    ADS  Google Scholar 

    44.
    Dingemanse, N. J. & Kalkman, V. J. Changing temperature regimes have advanced the phenology of Odonata in the Netherlands. Ecol. Ent. 33, 394–402 (2008).
    Google Scholar 

    45.
    McCauley, S. J., Hammond, J. I. & Mabry, K. E. Simulated climate change increases larval mortality, alters phenology, and affects flight morphology of a dragonfly. Ecosphere 9, e02151. https://doi.org/10.1002/ecs2.2151 (2018).
    Article  PubMed  PubMed Central  Google Scholar 

    46.
    Fincke, O. M. & Hadrys, H. Unpredictable offspring survivorship in the damselfly Megaloprepus coerulatus, shapes parental behavior, constraints sexual selection, and challenges traditional fitness-estimates. Evolution 55, 762–772 (2001).
    CAS  PubMed  Google Scholar 

    47.
    Johansson, F. Intraguild predation and cannibalism in odonate larvae: effects of foraging behaviour and zooplankton availability. Oikos 66, 80–87 (1993).
    Google Scholar 

    48.
    Parmesan, C. Ecological and evolutionary responses to recent climate change. Ann. Rev. Ecol. Evol. Syst. 37, 637–669 (2006).
    Google Scholar 

    49.
    SMHI, Swedish Meteorological and Hydrological Institute. Swedish air temperature, snow and wind; https://www.smhi.se/klimatdata (2017).

    50.
    Johansson, F. The slow–fast life style characteristics in a suite of six species of odonate larvae. Freshw. Biol. 43, 149–159 (2000).
    Google Scholar 

    51.
    Rychła, A., Benndorf, J. & Buczyński, P. Impact of pH and conductivity on species richness and community structure of dragonflies (Odonata) in small mining lakes. Fundam. Appl. Limnol. Arch. Hydrobiol. 179, 41–50 (2011).
    Google Scholar 

    52.
    National Register of Survey test-fishing – NORS. Swedish test fishing database; https://www.slu.se/en/departments/aquatic-resources1/databases1/national-register-of-survey-test-fishing-nors/ (2020).

    53.
    Robert, P.-A. Les Libellules (Delachaux & Niestlié, 1958).

    54.
    Harrington, R., Fleming, R. A. & Woiwod, I. P. Climate change impacts on insect management and conservation in temperate regions: Can they be predicted?. Agric. For. Entomol. 3, 233–240 (2001).
    Google Scholar 

    55.
    Bale, J. S. & Hayward, S. A. L. Insect overwintering in a changing climate. J. Exp. Biol. 213, 980–994 (2010).
    CAS  PubMed  Google Scholar 

    56.
    Grewe, Y., Hof, C., Dehling, D. M., Brandl, R. & Brändle, M. Recent range shifts of European dragonflies provide support for an inverse relationship between habitat predictability and dispersal. Glob. Ecol. Biogeogr. 22, 403–409 (2013).
    Google Scholar 

    57.
    Kalkman, V.J. et al. European Red List of Dragonflies. Publications Office of the European Union; https://ec.europa.eu/environment/nature/conservation/species/redlist/downloads/European_dragonflies.pdf (2010).

    58.
    Lande, R. Risks of population extinction from demographic and environmental stochasticity and random catastrophes. Am. Nat. 142, 911–927 (1993).
    PubMed  Google Scholar 

    59.
    Tsimring, L. S. Noise in biology. Rep. Prog. Phys. 77, 026601. https://doi.org/10.1088/0034-4885/77/2/026601 (2014).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    60.
    Popova, O. N., Haritonov, AYu. & Erdakov, L. N. Cyclicity of long-term population dynamics in dragonflies of the genus Sympetrum (Odonata, Anisoptera) in the basin of Lake Chany. Contemp. Probl. Ecol. 11, 551–562 (2018).
    Google Scholar 

    61.
    Sahlén, G. & Ekestubbe, K. Identification of dragonflies (Odonata) as indicators of general species richness in boreal forest lakes. Biodiv. Conserv. 10, 673–690 (2001).
    Google Scholar 

    62.
    Korkeamäki, E., Elo, M., Sahlén, G., Salmela, J. & Suhonen, J. Regional variations in occupancy frequency distribution patterns between odonate assemblages in Fennoscandia. Ecosphere 9, e02192. https://doi.org/10.1002/ecs2.2192 (2018).
    Article  Google Scholar 

    63.
    Angeler, D. G. & Johnson, R. K. Patterns of temporal community turnover are spatially synchronous across boreal lakes. Freshw. Biol. 57, 1782–1793 (2012).
    Google Scholar 

    64.
    MAGIC biblioteket sjöar. Lake data from Sweden; https://magicbiblioteket.ivl.se/ (2016).

    65.
    Swedish Forest Agency. Silvicultural activities; Planted area and Pre-commercially thinned area: 3-year average, 1000 hectares by region, year and ownership class; https://pxweb.skogsstyrelsen.se/pxweb/en/Skogsstyrelsens%20statistikdatabas/ (2016).

    66.
    SCB, Statistics Sweden. Land use in Sweden – Land use: Arable land and forest land by region and land use category; https://www.statistikdatabasen.scb.se/pxweb/en/ssd/START__MI__MI0803__MI0803A/MarkanvJbSk/ (2019).

    67.
    Hickling, R., Roy, D. B., Hill, J. K. & Thomas, C. D. A northward shift of range margins in British Odonata. Global Change Biol. 11, 502–506 (2005).
    ADS  Google Scholar 

    68.
    Conti, L., Schmidt-Kloiber, A., Grenouillet, G. & Graf, W. A trait-based approach to assess the vulnerability of European aquatic insects to climate change. Hydrobiologia 721, 297–315 (2014).
    CAS  Google Scholar 

    69.
    Lavorel, S. et al. Assessing functional diversity in the field–methodology matters!. Funct. Ecol. 22, 134–147 (2008).
    Google Scholar 

    70.
    Laliberté, E., Legendre, P. & Shipley, B. FD: measuring functional diversity from multiple traits, and other tools for functional ecology. R package version 1.0–12; https://cran.r-project.org/web/packages/FD/index.html (2014).

    71.
    R Development Core Team. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna. https://www.R-project.org (2019).

    72.
    Violle, C. et al. Let the concept of trait be functional!. Oikos 116, 882–892 (2007).
    Google Scholar 

    73.
    Dijkstra, K. D. B. & Lewington, R. Field Guide to the Dragonflies of Britain and Europe (British Wildlife Publishing, Devon, 2006).
    Google Scholar 

    74.
    Norling, U. Livscykler hos svenska Odonater. Entomologen 4, 1–14 (1975).
    Google Scholar 

    75.
    Norling, U. The life cycle and larval photoperiodic responses of Coenagrion hastulatum (Charpentier) in two climatically different areas (Zygoptera: Coenagrionidae). Odonatologica 13, 429–449 (1984).
    Google Scholar 

    76.
    Norling, U. Photoperiodic control of larval development in Leucorrhinia dubia (Vander Linden): a comparison between populations from northern and southern Sweden (Anisoptera: Libellulidae). Odonatologica 13, 529–550 (1984).
    Google Scholar  More

  • in

    Over 90 endangered fish and invertebrates are caught in industrial fisheries

    1.
    FAO. The State of World Fisheries and Aquaculture 2020: Sustainability in Action (FAO, 2020).
    2.
    Diaz, S. et al. 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 (2019).

    3.
    FAO. The State of World Fisheries and Aquaculture 2018—Meeting the Sustainable Development Goals (FAO, 2018).

    4.
    IUCN. The IUCN Red List of Threatened Species. Version 2019-1 (IUCN, 2019). https://www.iucnredlist.org.

    5.
    McClenachan, L., Cooper, A. B., Carpenter, K. E. & Dulvy, N. K. Extinction risk and bottlenecks in the conservation of charismatic marine species. Conserv. Lett. 5, 73–80 (2012).
    Google Scholar 

    6.
    McClenachan, L., Cooper, A. B. & Dulvy, N. K. Rethinking trade-driven extinction risk in marine and terrestrial Megafauna. Curr. Biol. 26, 1640–1646 (2016).
    CAS  PubMed  Google Scholar 

    7.
    Ripple, W. J. et al. Are we eating the world’s megafauna to extinction? Conserv. Lett. 12, e12627 (2019).
    Google Scholar 

    8.
    Di Minin, E., Leader-Williams, N. & Bradshaw, C. J. A. Banning trophy hunting will exacerbate biodiversity loss. Trends Ecol. Evol. 31, 99–102 (2016).
    PubMed  Google Scholar 

    9.
    Batavia, C. et al. The elephant (head) in the room: a critical look at trophy hunting. Conserv. Lett. 12, 1–6 (2019).
    ADS  Google Scholar 

    10.
    Nelson, M. P., Bruskotter, J. T., Vucetich, J. A. & Chapron, G. Emotions and the ethics of consequence in conservation decisions: lessons from cecil the lion. Conserv. Lett. 9, 302–306 (2016).
    Google Scholar 

    11.
    Aron, W., Burke, W. & Freeman, M. M. R. The whaling issue. Mar. Pol. 24, 179–191 (2000).
    Google Scholar 

    12.
    Shiffman, D. S., Gallagher, A. J., Wester, J., Macdonald, C. C. & Thaler, A. D. Trophy fishing for species threatened with extinction: a way forward building on a history of conservation. Mar. Pol. 50, 318–322 (2014).
    Google Scholar 

    13.
    Dulvy, N. K., Sadovy, Y. & Reynolds, J. D. Extinction vulnerability in marine populations. Fish Fish. 4, 25–64 (2003).
    Google Scholar 

    14.
    Pinsky, M. L., Jensen, O. P., Ricard, D. & Palumbi, S. R. Unexpected patterns of fisheries collapse in the world’s oceans. Proc. Natl Acad. Sci. USA 108, 8317–8322 (2011).
    ADS  CAS  PubMed  Google Scholar 

    15.
    Sumaila, U. R. et al. Updated estimates and analysis of global fisheries subsidies. Mar. Pol. 109, 103695 (2019).
    Google Scholar 

    16.
    Vincent, A. C. J., Sadovy de Mitcheson, Y. J., Fowler, S. L. & Lieberman, S. The role of CITES in the conservation of marine fishes subject to international trade. Fish Fish. 15, 563–592 (2014).
    Google Scholar 

    17.
    Crespo, G. O. et al. High-seas fish biodiversity is slipping through the governance net. Nat. Ecol. Evol. 3, 1273–1276 (2019).
    PubMed  Google Scholar 

    18.
    Reiss, H., Hoarau, G., Dickey-Collas, M. & Wolff, W. J. Genetic population structure of marine fish: mismatch between biological and fisheries management units. Fish Fish. 10, 361–395 (2009).
    Google Scholar 

    19.
    Collette, B. B. et al. High value and long life—double jeopardy for tunas and billfi shes. Science 333, 291–292 (2011).
    ADS  CAS  PubMed  Google Scholar 

    20.
    Neubauer, P., Jensen, O. P., Hutchings, J. A. & Baum, J. K. Resilience and recovery of overexploited marine populations. Science 340, 347–349 (2013).
    ADS  CAS  PubMed  Google Scholar 

    21.
    Hutchings, J. A. Collapse and recovery of marine fishes. Nature 406, 882–885 (2000).
    ADS  CAS  PubMed  Google Scholar 

    22.
    Doney, S. C. et al. Climate change impacts on marine ecosystems. Annu. Rev. Mar. Sci. 4, 11–37 (2012).
    ADS  Google Scholar 

    23.
    Ye, Y. & Gutierrez, N. L. Ending fishery overexploitation by expanding from local successes to globalized solutions. Nat. Ecol. Evol. 1, 0179 (2017).
    Google Scholar 

    24.
    Williams, B. R., Burgess, M. G., Ashe, E., Gaines, S. D. & Randall, R. Marine mammal protections require increased global capacity. Science 354, 1372–1374 (2016).
    ADS  CAS  PubMed  Google Scholar 

    25.
    Watson, R. & Pauly, D. Systematic distortions in world fisheries catch trends. Nature 414, 534–536 (2001).
    ADS  CAS  PubMed  Google Scholar 

    26.
    Powles, H. et al. Assessing and protecting endangered marine species. ICES J. Mar. Sci. 57, 669–676 (2000).
    Google Scholar 

    27.
    Dent, F. & Clarke, S. State of the Global Market for Shark Products. FAO Fisheries and Aquaculture Technical Paper no. 590 (Food and Agriculture Organization of the United Nations, Rome 2015).

    28.
    Oliver, S., Braccini, M., Newman, S. J. & Harvey, E. S. Global patterns in the bycatch of sharks and rays. Mar. Pol. 54, 86–97 (2015).
    Google Scholar 

    29.
    Davies, R. W. D., Cripps, S. J., Nickson, A. & Porter, G. Defining and estimating global marine fisheries bycatch. Mar. Pol. 33, 661–672 (2009).
    Google Scholar 

    30.
    Watson, R. A., Green, B. S., Tracey, S. R., Farmery, A. & Pitcher, T. J. Provenance of global seafood. Fish Fish. 17, 585–595 (2016).
    Google Scholar 

    31.
    Jabado, R. W. et al. Troubled waters: threats and extinction risk of the sharks, rays and chimaeras of the Arabian Sea and adjacent waters. Fish Fish. 19, 1043–1062 (2018).
    Google Scholar 

    32.
    Anticamara, J. A., Watson, R., Gelchu, A. & Pauly, D. Global fishing effort (1950-2010): trends, gaps, and implications. Fish. Res. 107, 131–136 (2011).
    Google Scholar 

    33.
    Anderson, S. C., Flemming, J. M., Watson, R. & Lotze, H. K. Serial exploitation of global sea cucumber fisheries. Fish Fish. 12, 317–339 (2011).
    Google Scholar 

    34.
    Webb, T. J. & Mindel, B. L. Global patterns of extinction risk in marine and non-marine systems. Curr. Biol. 25, 506–511 (2015).
    CAS  PubMed  Google Scholar 

    35.
    Dulvy, et al. Extinction risk and conservation of the world’s sharks and rays. Elife 3, e00590 (2014).
    PubMed  PubMed Central  Google Scholar 

    36.
    Anderson, S. C., Flemming, J. M., Watson, R. & Lotze, H. K. Rapid global expansion of invertebrate fisheries: trends, drivers, and ecosystem effects. PLoS ONE 6, 1–9 (2011).
    Google Scholar 

    37.
    Willette, D. A. & Cheng, S. H. Delivering on seafood traceability under the new U.S. import monitoring program. AMBIO 47, 25–30 (2018).
    PubMed  Google Scholar 

    38.
    Simpfendorfer, C. A. & Dulvy, N. K. Bright spots of sustainable shark fishing. Curr. Biol. 27, R97–R98 (2017).
    CAS  PubMed  Google Scholar 

    39.
    Hobbs, C. A. D. et al. Using DNA barcoding to investigate patterns of species utilisation in UK shark products reveals threatened species on sale. Sci. Rep. 9, 1–10 (2019).
    Google Scholar 

    40.
    Porszt, E. J., Peterman, R. M., Dulvy, N. K., Cooper, A. B. & Irvine, J. R. Reliability of indicators of decline in abundance. Conserv. Biol. 26, 894–904 (2012).
    PubMed  Google Scholar 

    41.
    D’Eon-Eggertson, F., Dulvy, N. K. & Peterman, R. M. Reliable identification of declining populations in an uncertain world. Conserv. Lett. 8, 86–96 (2015).
    Google Scholar 

    42.
    Foley, C. M., Lynch, M. A., Thorne, L. H. & Lynch, H. J. Listing foreign species under the endangered species act: a primer for conservation biologists. Bioscience 67, 627–637 (2017).
    Google Scholar 

    43.
    Sky, M. B. Getting on the list: politics and procedural maneuvering in cites appendix I and II decisions for commercially exploited marine and timber species. Sustain. Dev. Law Policy 10, 35–40 (2010).
    Google Scholar 

    44.
    UNEP‐WCMC (2019). The checklist of CITES species website. http://checklist.cites.org.

    45.
    Probst, W. N. How emerging data technologies can increase trust and transparency in fisheries. ICES J. Mar. Sci. 77, 1286–1294 (2019).
    Google Scholar 

    46.
    Lewis, S. G. & Boyle, M. The expanding role of traceability in seafood: tools and key initiatives. J. Food Sci. 82, A13–A21 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    47.
    Kamilaris, A., Fonts, A. & Prenafeta-Boldύ, F. X. The rise of blockchain technology in agriculture and food supply chains. Trends Food Sci. Technol. 91, 640–652 (2019).
    CAS  Google Scholar 

    48.
    WWF. WWF-Australia and OpenSc (Panda Labs, 2019). https://www.wwf.org.au/get-involved/panda-labs/opensc#gs.hi7rtb.

    49.
    Boulais, O. Exploring Provenance of Tuna using Distributed Ledgers. (Viral Communications, 2019). https://viral.media.mit.edu/pub/tunaprovenance.

    50.
    Hosch, G. & Blaha, F. Seafood Traceability for Fisheries Compliance: Country-Level Support for Catch Documentation Schemes. FAO Fisheries and Aquaculture Technical Paper 619 (Food and Agriculture Organization of the United Nations, Rome, 2017).

    51.
    Miller, N. A., Roan, A., Hochberg, T., Amos, J. & Kroodsma, D. A. Identifying global patterns of transshipment behavior. Front. Mar. Sci. 5, 1–9 (2018).
    Google Scholar 

    52.
    Miller, A. M. M., Bush, S. R. & Mol, A. P. J. Power Europe: EU and the illegal, unreported and unregulated tuna fisheries regulation in the West and Central Pacific Ocean. Mar. Pol. https://doi.org/10.1016/j.marpol.2013.12.009 (2014).

    53.
    Osterblom, H. et al. Transnational corporations as ‘keystone actors’ in marine ecosystems. PLoS ONE 10, 1–15 (2015).
    Google Scholar 

    54.
    Davies, T. D. & Baum, J. K. Extinction risk and overfishing: reconciling conservation and fisheries perspectives on the status of marine fishes. Sci. Rep. 2, 561 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    55.
    Hornborg, S., Svensson, M., Nilsson, P. & Ziegler, F. By-catch impacts in fisheries: utilizing the IUCN red list categories for enhanced product level assessment in seafood LCAS. Environ. Manag. 52, 1239–1248 (2013).
    ADS  Google Scholar 

    56.
    Dulvy, N. K., Jennings, S., Goodwin, N. B., Grant, A. & Reynolds, J. D. Comparison of threat and exploitation status in North-East Atlantic marine populations. J. Appl. Ecol. 42, 883–891 (2005).
    Google Scholar 

    57.
    Mace, G. M. et al. Quantification of extinction risk: IUCN’s system for classifying threatened species. Conserv. Biol. 22, 1424–1442 (2008).
    PubMed  Google Scholar 

    58.
    Fernandes et al. Coherent assessments of Europe’s marine fishes show regional divergence and megafauna loss. Nat. Ecol. Evol. 1, 0170 (2017).
    Google Scholar 

    59.
    Pauly, D., Zeller, D. & Palomares, M. L. D. Sea Around us Concepts, Design and Data. seaaroundus.org (University of British Columbia, Vancouver, BC 2020).

    60.
    Watson, R. A. & Tidd, A. Mapping nearly a century and a half of global marine fishing: 1869–2015. Mar. Pol. 93, 171–177 (2018).
    Google Scholar 

    61.
    Pauly, D. & Zeller, D. Agreeing with FAO: comments on SOFIA 2018. Mar. Pol. 100, 332–333 (2019).
    Google Scholar 

    62.
    Tai, T. C., Cashion, T., Lam, V. W. Y., Swartz, W. & Sumaila, U. R. Ex-vessel fish price database: disaggregating prices for low-priced species from reduction fisheries. Front. Mar. Sci. 4, 1–10 (2017).
    Google Scholar  More