More stories

  • in

    Altitude and hillside orientation shapes the population structure of the Leishmania infantum vector Phlebotomus ariasi

    1.
    Dolmatova, A. V. & Demina, N. A. Les phlébotomes (Phlebotominae) et les maladies qu’ils transmettent. ORSTOM 20, 1–169 (1966).
    Google Scholar 
    2.
    Bichaud, L. et al. Epidemiologic relationship between toscana virus infection and Leishmania infantum due to common exposure to Phlebotomus perniciosus sandfly vector. PLoS Negl. Trop. Dis. 5(9), e1328 (2011).
    PubMed  PubMed Central  Google Scholar 

    3.
    Rioux, J.-A., Killick-Kendrick, R., Perieres, J., Turner, D. & Lanotte, G. Ecologie des Leishmanioses dans le sud de la France. 13. Les sites de “flanc de coteau”, biotopes de transmission privilégiés de la Leishmaniose viscérale en Cévennes. Ann . Parasitol. Hum. Comp. 55(4), 445–453 (1980).
    CAS  PubMed  Google Scholar 

    4.
    Rioux, J.-A. et al. Ecology of leishmaniasis in the South of France. 22. Reliability and representativeness of 12 Phlebotomus ariasi, P. perniciosus and Sergentomyia minuta (Diptera: Psychodidae) sampling stations in Vallespir (eastern French Pyrenees region). Parasite 20, 34 (2013).
    PubMed  PubMed Central  Google Scholar 

    5.
    Rioux, J.-A. et al. Epidémiologie des leishmanioses dans le Sud de la France. Monogr. l’Inst. Natl. Santé Rech. Méd. 20, 1–228 (1969).
    Google Scholar 

    6.
    Lewis, D. J. A taxonomic review of the genus Phlebotomus (Diptera: Psychodidae). Bull. Br. Museum 45(2), 121–209 (1982).
    Google Scholar 

    7.
    Rossi, E. et al. Mapping the main Leishmania phlebotomine vector in the endemic focus of the Mt. Vesuvius in southern Italy. Geospat. Health 1(2), 191–198 (2007).
    PubMed  Google Scholar 

    8.
    Ballart, C., Barón, S., Alcover, M. M., Portus, M. & Gallego, M. Distribution of phlebotomine sand flies (Diptera: Psychodidae) in Andorra: First finding of P. perniciosus and wide distribution of P. ariasi. Acta Trop. 122(1), 155–159 (2012).
    CAS  PubMed  Google Scholar 

    9.
    Ballart, C. et al. Importance of individual analysis of environmental and climatic factors affecting the density of Leishmania vectors living in the same geographical area: The example of Phlebotomus ariasi and P. perniciosus in northeast Spain. Geospat. Health 8(2), 389–403 (2014).
    PubMed  Google Scholar 

    10.
    Boussaa, S., Neffa, M., Pesson, B. & Boumezzough, A. Phlebotomine sandflies (Diptera: Psychodidae) of southern Morocco: Results of entomological surveys along the Marrakech-Ouarzazat and Marrakech-Azilal roads. Ann. Trop. Med. Parasitol. 104(2), 163–170 (2010).
    CAS  PubMed  Google Scholar 

    11.
    Franco, F. et al. Genetic structure of Phlebotomus (Larroussius) ariasi populations, the vector of Leishmania infantum in the western Mediterranean: Epidemiological implications. Int. J. Parasitol. 40(11), 1335–1346 (2010).
    PubMed  Google Scholar 

    12.
    Ready, P. Leishmaniasis emergence in Europe. Euro Surveill. 15(10), 19505 (2010).
    CAS  PubMed  Google Scholar 

    13.
    Branco, S. et al. Entomological and ecological studies in a new potential zoonotic leishmaniasis focus in Torres Novas municipality, Central Region, Portugal. Acta Trop. 125(3), 339–348 (2013).
    CAS  PubMed  Google Scholar 

    14.
    Barón, S. D. et al. Risk maps for the presence and absence of Phlebotomus perniciosus in an endemic area of leishmaniasis in southern Spain: Implications for the control of the disease. Parasitology 138(10), 1234–1244 (2011).
    PubMed  Google Scholar 

    15.
    Boudabous, R. et al. The phlebotomine fauna (Diptera: Psychodidae) of the eastern coast of Tunisia. J. Med. Entomol. 46(1), 1–8 (2009).
    CAS  PubMed  Google Scholar 

    16.
    European Centre for Disease Prevention and Control E. Phlebotomine sand flies maps [internet] 2019 [10/01/19]. https://www.ecdc.europa.eu/en/disease-vectors/surveillance-and-disease-data/phlebotomine-maps.

    17.
    Dedet, J.-P. Les leishmanioses en France métropolitaine. BEH Hors-Sér. 2010, 9–12 (2020).
    Google Scholar 

    18.
    Depaquit, J., Grandadam, M., Fouque, F., Andry, P.-E. & Peyrefitte, C. Arthropod-borne viruses transmitted by Phlebotomine sandflies in Europe: A review. Euro Surveill. 15(10), 19507 (2010).
    CAS  PubMed  Google Scholar 

    19.
    Kamhawi, S. et al. Two populations of Phlebotomus ariasi in the Cévennes focus of leishmaniasis in the south of France revealed by analysis of cuticular hydrocarbons. Med. Vet. Entomol. 1(1), 97–102 (1987).
    CAS  PubMed  Google Scholar 

    20.
    Pesson, B., Wallon, M., Floer, M. & Kristensen, A. Étude isoenzymatique de populations méditerranéennes de phlébotomes du sous-genre Larroussius. Parassitologia 33, 471–476 (1991).
    PubMed  Google Scholar 

    21.
    Ballart, C., Pesson, B. & Gallego, M. Isoenzymatic characterization of Phlebotomus ariasi and P. perniciosus of canine leishmaniasis foci from Eastern Pyrenean regions and comparison with other populations from Europe. Parasite. 25, 3 (2018).
    PubMed  PubMed Central  Google Scholar 

    22.
    Martin-Sanchez, J., Gramiccia, M., Pesson, B. & Morillas-Marquez, F. Genetic polymorphism in sympatric species of the genus Phlebotomus, with special reference to Phlebotomus perniciosus and Phlebotomus longicuspis (Diptera, Phlebotomidae). Parasite 7(4), 247–254 (2000).
    CAS  PubMed  Google Scholar 

    23.
    Mahamdallie, S. S., Pesson, B. & Ready, P. D. Multiple genetic divergences and population expansions of a Mediterranean sandfly, Phlebotomus ariasi, in Europe during the Pleistocene glacial cycles. Heredity 106(5), 714–726 (2010).
    PubMed  PubMed Central  Google Scholar 

    24.
    Prudhomme, J. et al. Ecology and spatiotemporal dynamics of sandflies in the Mediterranean Languedoc region (Roquedur area, Gard, France). Parasit. Vectors 8(1), 1–14 (2015).
    Google Scholar 

    25.
    Prudhomme, J. et al. Ecology and morphological variations in wings of Phlebotomus ariasi (Diptera: Psychodidae) in the region of Roquedur (Gard, France): A geometric morphometrics approach. Parasit. Vectors 9(1), 578 (2016).
    PubMed  PubMed Central  Google Scholar 

    26.
    Lachaud, L. et al. Surveillance of leishmaniases in France, 1999 to 2012. Euro Surveill. 18(29), 20534 (2013).
    CAS  PubMed  Google Scholar 

    27.
    Prudhomme, J. et al. New microsatellite markers for multi-scale genetic studies on Phlebotomus ariasi Tonnoir, vector of Leishmania infantum in the Mediterranean area. Acta Trop. 142, 79–85 (2015).
    CAS  PubMed  Google Scholar 

    28.
    Wattier, R., Engel, C. R., Saumitou-Laprade, P. & Valero, M. Short allele dominance as a source of heterozygote deficiency at microsatellite loci: Experimental evidence at the dinucleotide locus Gv1CT in Gracilaria gracilis (Rhodophyta). Mol. Ecol. 7(11), 1569–1573 (1998).
    CAS  Google Scholar 

    29.
    De Meeûs T, Chan CT, Ludwig JM, Tsao JI, Patel J, Bhagatwala J, Beati L. Deceptive combined effects of short allele dominance and stuttering: An example with Ixodes scapularis, the main vector of Lyme disease in the U.S.A. peerreviewed and recommended by PCI Evolutionary Biology. 2019.

    30.
    De Meeûs, T. Revisiting, FIS, FST, Wahlund Effects, and Null Alleles. J. Hered. 109(4), 446–456 (2018).
    PubMed  Google Scholar 

    31.
    Teriokhin, A. T., De Meeûs, T. & Guegan, J. F. On the power of some binomial modifications of the Bonferroni multiple test. J. Gener. Biol. 68(5), 332–340 (2007).
    CAS  Google Scholar 

    32.
    De Meeûs, T., Guégan, J.-F. & Teriokhin, A. T. MultiTest V.1.2., a program to binomially combine independent tests and performance comparison with other related methods on proportional data. BMC Bioinform. 10(1), 443 (2009).
    Google Scholar 

    33.
    Nei, M. & Chesser, R. K. Estimation of fixation indices and gene diversities. Ann. Hum. Genet. 47(3), 253–259 (1983).
    CAS  PubMed  MATH  Google Scholar 

    34.
    Meirmans, P. G. & Hedrick, P. W. Assessing population structure: FST and related measures. Mol. Ecol. Resour. 11(1), 5–18 (2011).
    PubMed  Google Scholar 

    35.
    Wang, J. Does GST underestimate genetic differentiation from marker data? Mol. Ecol. 24(14), 3546–3558 (2015).
    CAS  PubMed  Google Scholar 

    36.
    Chapuis, M. P. & Estoup, A. Microsatellite null alleles and estimation of population differentiation. Mol. Biol. Evol. 24(3), 621–631 (2007).
    CAS  PubMed  Google Scholar 

    37.
    Maingon, R. et al. Genetic identification of two sibling species of Lutzomyia longipalpis (Diptera: Psychodidae) that produce distinct male sex pheromones in Sobral, Ceará State, Brazil. Mol. Ecol. 12(7), 1879–1894 (2003).
    CAS  PubMed  Google Scholar 

    38.
    Bauzer, L. G., Souza, N. A., Maingon, R. D. & Peixoto, A. A. Lutzomyia longipalpis in Brazil: A complex or a single species? A mini-review. Mem. Inst. Oswaldo Cruz. 102(1), 1–12 (2007).
    PubMed  Google Scholar 

    39.
    Scarpassa, V. M. & Alencar, R. B. Lutzomyia umbratilis, the main vector of Leishmania guyanensis, represents a novel species complex? PLoS One 7(5), e37341 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    40.
    Tharmatha, T., Gajapathy, K., Ramasamy, R. & Surendran, S. N. Morphological and molecular identification of cryptic species in the Sergentomyia bailyi (Sinton, 1931) complex in Sri Lanka. Bull. Entomol. Res. 107(1), 58–65 (2016).
    PubMed  Google Scholar 

    41.
    Balloux, F. Heterozygote excess in small populations and the heterozygote-excess effective population size. Evolution 58(9), 1891–1900 (2004).
    PubMed  Google Scholar 

    42.
    Manangwa, O. et al. Detecting Wahlund effects together with amplification problems: Cryptic species, null alleles and short allele dominance in Glossina pallidipes populations from Tanzania. Mol. Ecol. Resour. 19(3), 757–772 (2019).
    PubMed  Google Scholar 

    43.
    Hartl, D. L. & Clarck, A. G. Principles of Population Genetics 2nd edn. (Sinauer Associates Inc, Sunderland, 1989).
    Google Scholar 

    44.
    Araki, A. S. et al. Multilocus analysis of divergence and introgression in sympatric and allopatric sibling species of the Lutzomyia longipalpis complex in Brazil. PLoS Negl Trop Dis. 7(10), e2495 (2013).
    PubMed  PubMed Central  Google Scholar 

    45.
    Kyriacou, C. Sex and rhythms in sandflies and mosquitoes: an appreciation of the work of Alexandre Afranio Peixoto (1963–2013). Infect. Genet. Evol. 28, 662–665 (2014).
    PubMed  PubMed Central  Google Scholar 

    46.
    Abonnenc E. Les phlébotomes de la région éthiopienne (Diptera, Psychodidae): Cahiers de l’ORSTOM, série Entomologie médicale et Parasitologie; 1972 01/01. 239.

    47.
    Rougeron, V. et al. Reproductive strategies and population structure in Leishmania: Substantial amount of sex in Leishmania Viannia guyanensis. Mol. Ecol. 20(15), 3116–3127 (2011).
    PubMed  Google Scholar 

    48.
    Rougeron, V. et al. Multifaceted population structure and reproductive strategy in Leishmania donovani complex in one Sudanese village. PLoS Negl. Trop. Dis. 5(12), e1448 (2011).
    PubMed  PubMed Central  Google Scholar 

    49.
    Rioux, J.-A. et al. Ecologie des Leishmanioses dans le sud de la France. 12. Dispersion horizontale de Phlebotomus ariasi Tonnoir, 1921. Experiences préliminaires. Ann. Parasitol. Hum. Comp. 54(6), 673–682 (1979).
    CAS  PubMed  Google Scholar 

    50.
    Latch, E. K., Dharmarajan, G., Glaubitz, J. C. & Rhodes, O. E. Relative performance of Bayesian clustering software for inferring population substructure and individual assignment at low levels of population differentiation. Conserv. Genet. 7(2), 295–302 (2006).
    Google Scholar 

    51.
    Kaeuffer, R., Réale, D., Coltman, D. & Pontier, D. Detecting population structure using STRUCTURE software: Effect of background linkage disequilibrium. Heredity 99(4), 374–380 (2007).
    CAS  PubMed  Google Scholar 

    52.
    Frantz, A. C., Cellina, S., Krier, A., Schley, L. & Burke, T. Using spatial Bayesian methods to determine the genetic structure of a continuously distributed population: Clusters or isolation by distance? J. Appl. Ecol. 46(2), 493–505 (2009).
    Google Scholar 

    53.
    Blair, C. et al. A simulation-based evaluation of methods for inferring linear barriers to gene flow. Mol. Ecol. Resour. 12(5), 822–833 (2012).
    PubMed  Google Scholar 

    54.
    Bohling, J. H. et al. Describing a developing hybrid zone between red wolves and coyotes in eastern North Carolina, USA. Evol. Appl. 9(6), 791–804 (2016).
    PubMed  PubMed Central  Google Scholar 

    55.
    Le, D. P. bioclimat Mediterraneen: Analyse des formes climatiques par le systeme d’Emberger. Vegetation 34(2), 87–103 (1977).
    Google Scholar 

    56.
    Alten, B. et al. Sampling strategies for phlebotomine sand flies (Diptera: Psychodidae) in Europe. Bull. Entomol. Res. 105(6), 664–678 (2015).
    CAS  PubMed  Google Scholar 

    57.
    Ayhan, N. et al. Practical guidelines for studies on sandfly-borne phleboviruses: Part I: Important points to consider ante field work. Vector Borne Zoonot. Dis. 17(1), 73–80 (2017).
    Google Scholar 

    58.
    Killick-Kendrick, R. et al. The identification of female sandflies of the subgenus Larroussius by the morphology of the spermathecal ducts. Parassitologia 33, 335–347 (1991).
    PubMed  Google Scholar 

    59.
    Wang, Q. & Wang, X. Comparison of methods for DNA extraction from a single chironomid for PCR analysis. Pak. J. Zool. 44(2), 421–426 (2012).
    CAS  Google Scholar 

    60.
    Esseghir, S., Ready, P. D., Killick-Kendrick, R. & Ben-Ismail, R. Mitochondrial haplotypes and phylogeography of Phlebotomus vectors of Leishmania major. Insect. Mol. Biol. 6(3), 221–225 (1997).
    Google Scholar 

    61.
    Depaquit, J., Leger, N. & Randrianambinintsoa, F. J. Paraphyly of the subgenus Anaphlebotomus and creation of Madaphlebotomus subg. Nov. (Phlebotominae: Phlebotomus). Med. Vet. Entomol. 29(2), 159–170 (2015).
    CAS  PubMed  Google Scholar 

    62.
    Coombs, J. A., Letcher, B. H. & Nislow, K. H. Create: A software to create input files from diploid genotypic data for 52 genetic software programs. Mol. Ecol. Resour. 8(3), 578–580 (2008).
    CAS  PubMed  Google Scholar 

    63.
    Bohling, J. H., Adams, J. R. & Waits, L. P. Evaluating the ability of Bayesian clustering methods to detect hybridization and introgression using an empirical red wolf data set. Mol. Ecol. 22(1), 74–86 (2013).
    PubMed  Google Scholar 

    64.
    Jombart, T., Devillard, S. & Balloux, F. Discriminant analysis of principal components: A new method for the analysis of genetically structured populations. BMC Genet. 11(1), 94 (2010).
    PubMed  PubMed Central  Google Scholar 

    65.
    Jombart, T. adegenet: A R package for the multivariate analysis of genetic markers. Bioinformatics 24(11), 1403–1405 (2008).
    CAS  Google Scholar 

    66.
    R Development Core Team RT. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria, 2012. https://www.R-project.org/. 2018.

    67.
    Corander, J. & Marttinen, P. Bayesian identification of admixture events using multilocus molecular markers. Mol. Ecol. 15(10), 2833–2843 (2006).
    PubMed  Google Scholar 

    68.
    Corander, J., Marttinen, P., Siren, J. & Tang, J. Enhanced Bayesian modelling in BAPS software for learning genetic structures of populations. BMC Bioinform. 9, 539 (2008).
    Google Scholar 

    69.
    Pritchard, J. K., Stephens, M. & Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 155(2), 945–959 (2000).
    CAS  PubMed  PubMed Central  Google Scholar 

    70.
    Earl, D. A. & vonHoldt, B. M. STRUCTURE HARVESTER: A website and program for visualizing STRUCTURE output and implementing the Evanno method. Conserv. Genet. Resourc. 4(2), 359–361 (2012).
    Google Scholar 

    71.
    Kumar, S., Stecher, G. & Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33(7), 1870–1874 (2016).
    CAS  Google Scholar 

    72.
    Cavalli-Sforza, L. L. & Edwards, A. W. F. Phylogenetic analysis. Models and estimation procedures. Am. J. Hum. Genet. 19(3 Pt 1), 233–257 (1967).
    CAS  PubMed  PubMed Central  Google Scholar 

    73.
    Takezaki, N. & Nei, M. Genetic distances and reconstruction of phylogenetic trees from microsatellite DNA. Genetics 144(1), 389–399 (1996).
    CAS  PubMed  PubMed Central  Google Scholar 

    74.
    Goudet, J. FSTAT (Version 1.2): A computer program to calculate F-statistics. J. Hered. 86(6), 485–486 (1995).
    Google Scholar 

    75.
    Benjamini, Y. & Yekutieli, D. The control of the false discovery rate in multiple testing under dependency. Ann. Stat. 29(4), 1165–1188 (2001).
    MathSciNet  MATH  Google Scholar 

    76.
    Wright, S. The interpretation of population structure by F-statistics with special regard to systems of mating. Evolution 19(3), 395–420 (1965).
    Google Scholar 

    77.
    Weir, B. S. & Cockerham, C. C. Estimating F-statistics for the analysis of population structure. Evolution 38(6), 1358–1370 (1984).
    CAS  PubMed  Google Scholar 

    78.
    Goudet, J., Raymond, M., De Meeûs, T. & Rousset, F. Testing differentiation in diploid populations. Genetics 20, 144 (1996).
    Google Scholar 

    79.
    De Meeûs, T. et al. Population genetics and molecular epidemiology or how to “débusquer la bête”. Infect. Genet. Evol. 20, 7 (2007).
    Google Scholar 

    80.
    Van Oosterhout, C., Hutchinson, W. F., Wills, D. P. & Shipley, P. MICRO-CHECKER: Software for identifying and correcting genotyping errors in microsatellite data. Mol. Ecol. Notes. 4(3), 535–538 (2004).
    Google Scholar 

    81.
    Brookfield, J. F. A simple new method for estimating null allele frequency from heterozygote deficiency. Mol. Ecol. 5(3), 453–455 (1996).
    CAS  PubMed  Google Scholar 

    82.
    Frontier, S. Étude de la décroissance des valeurs propres dans une analyse en composantes principales: Comparaison avec le modèle du bâton brisé. J. Exp. Mar. Biol. Ecol. 25, 67–75 (1976).
    Google Scholar 

    83.
    Fox, J. & The, R. Commander: A basic-statistics graphical user interface to R. J. Stat. Softw. 14(9), 1–42 (2005).
    Google Scholar 

    84.
    Fox, J. Extending the R Commander by “Plug-In” Packages. R News 7(3), 46–52 (2007).
    Google Scholar 

    85.
    Akaïke, H. A new look at the statistical model identification. IEEE Trans. Autom. Control 19(6), 716–723 (1974).
    ADS  MathSciNet  MATH  Google Scholar 

    86.
    Rousset, F. GENEPOP’007: A complete re-implementation of the GENEPOP software for Windows and Linux. Mol. Ecol. Resour. 20, 8 (2008).
    Google Scholar 

    87.
    Séré, M., Thevenon, S., Belem, A. M. G. & De Meeus, T. Comparison of different genetic distances to test isolation by distance between populations. Heredity 119(2), 55–63 (2017).
    PubMed  PubMed Central  Google Scholar 

    88.
    Mantel, N. The detection of disease clustering and a generalized regression approach. Cancer Res. 27(2), 209–220 (1967).
    CAS  PubMed  Google Scholar 

    89.
    De Meeûs, T. Statistical decision from k test series with particular focus on population genetics tools: A DIY notice. Infect. Genet. Evol. 22, 91–93 (2014).
    PubMed  Google Scholar 

    90.
    Do, C. et al. NeEstimator v2: Re-implementation of software for the estimation of contemporary effective population size (Ne) from genetic data. Mol. Ecol. Resour. 14(1), 209–214 (2014).
    CAS  PubMed  Google Scholar 

    91.
    Waples, R. S. A bias correction for estimates of effective population size based on linkage disequilibrium at unlinked gene loci*. Conserv. Genet. 7(2), 167–184 (2006).
    Google Scholar 

    92.
    Waples, R. S. & Do, C. ldne: A program for estimating effective population size from data on linkage disequilibrium. Mol. Ecol. Resour. 8(4), 753–756 (2008).
    PubMed  Google Scholar 

    93.
    Peel, D., Waples, R. S., Macbeth, G. M., Do, C. & Ovenden, J. R. Accounting for missing data in the estimation of contemporary genetic effective population size (Ne). Mol. Ecol. Resour. 13(2), 243–253 (2013).
    CAS  PubMed  Google Scholar 

    94.
    Nomura, T. Estimation of effective number of breeders from molecular coancestry of single cohort sample. Evol. Appl. 1(3), 462–474 (2008).
    PubMed  PubMed Central  Google Scholar 

    95.
    Vitalis, R. & Couvet, D. Estimation of effective population size and migration rate from one- and two-locus identity measures. Genetics 157(2), 911–925 (2001).
    CAS  PubMed  PubMed Central  Google Scholar 

    96.
    Vitalis, R. & Couvet, D. Estim 1.0: A computer program to infer population parameters from one- and two-locus gene identity probabilities. Mol. Ecol. Notes 1(4), 354–356 (2005).
    Google Scholar  More

  • in

    Accelerating homogenization of the global plant–frugivore meta-network

    1.
    Helmus, M. R., Mahler, D. L. & Losos, J. B. Island biogeography of the Anthropocene. Nature 513, 543–546 (2014).
    ADS  CAS  PubMed  Google Scholar 
    2.
    Capinha, C., Essl, F., Seebens, H., Moser, D. & Pereira, H. M. Biogeography. The dispersal of alien species redefines biogeography in the Anthropocene. Science 348, 1248–1251 (2015).
    ADS  CAS  PubMed  Google Scholar 

    3.
    Seebens, H. et al. No saturation in the accumulation of alien species worldwide. Nat. Commun. 8, 14435 (2017).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    4.
    David, P. et al. Impacts of invasive species on food webs: a review of empirical data. Adv. Ecol. Res. 56, 1–60 (2017).
    Google Scholar 

    5.
    Traveset, A. & Richardson, D. M. Mutualistic interactions and biological invasions. Annu. Rev. Ecol. Evol. Syst. 45, 89–113 (2014).
    Google Scholar 

    6.
    Guimarães, P. R. Jr, Pires, M. M., Jordano, P., Bascompte, J. & Thompson, J. N. Indirect effects drive coevolution in mutualistic networks. Nature 550, 511–514 (2017).
    ADS  PubMed  Google Scholar 

    7.
    McKinney, M. L. & Lockwood, J. L. Biotic homogenization: a few winners replacing many losers in the next mass extinction. Trends Ecol. Evol. 14, 450–453 (1999).
    CAS  PubMed  Google Scholar 

    8.
    Wallace, A. R. The Geographical Distribution of Animals and Plants (Harper & Brothers, 1876).

    9.
    Holt, B. G. et al. An update of Wallace’s zoogeographic regions of the world. Science 339, 74–78 (2013).
    ADS  CAS  PubMed  Google Scholar 

    10.
    Vizentin-Bugoni, J. et al. Structure, spatial dynamics, and stability of novel seed dispersal mutualistic networks in Hawai’i. Science 364, 78–82 (2019).
    ADS  CAS  PubMed  Google Scholar 

    11.
    Jordano, P. in Seeds: The Ecology of Regeneration in Plant Communities 2nd edn (ed. Fenner, M.) 18–61 (CABI, 2000).

    12.
    Nathan, R. & Muller-Landau, H. C. Spatial patterns of seed dispersal, their determinants and consequences for recruitment. Trends Ecol. Evol. 15, 278–285 (2000).
    CAS  PubMed  Google Scholar 

    13.
    Herrera, C. M. in Plant–Animal Interactions: An Evolutionary Approach (eds Herrera, C. M. & Pellmyr, O.) 185–208 (Blackwell, 2002).

    14.
    Jordano, P. Patterns of mutualistic interactions in pollination and seed dispersal: connectance, dependence asymmetries, and coevolution. Am. Nat. 129, 657–677 (2002).
    Google Scholar 

    15.
    Tiffney, B. H. Vertebrate dispersal of seed plants through time. Annu. Rev. Ecol. Evol. Syst. 35, 1–29 (2004).
    Google Scholar 

    16.
    McConkey, K. R. et al. Seed dispersal in changing landscapes. Biol. Conserv. 146, 1–13 (2012).
    Google Scholar 

    17.
    Terborgh, J. et al. Tree recruitment in an empty forest. Ecology 89, 1757–1768 (2008).
    PubMed  Google Scholar 

    18.
    Galetti, M. et al. Functional extinction of birds drives rapid evolutionary changes in seed size. Science 340, 1086–1090 (2013).
    ADS  CAS  PubMed  Google Scholar 

    19.
    Rogers, H. S. et al. Effects of an invasive predator cascade to plants via mutualism disruption. Nat. Commun. 8, 14557 (2017).
    ADS  PubMed  PubMed Central  Google Scholar 

    20.
    Aslan, C. E., Zavaleta, E. S., Tershy, B. & Croll, D. Mutualism disruption threatens global plant biodiversity: a systematic review. PLoS One 8, e66993 (2013).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    21.
    Schleuning, M. et al. Ecological networks are more sensitive to plant than to animal extinction under climate change. Nat. Commun. 7, 13965 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    22.
    Thompson, J. The Geographic Mosaic of Coevolution (Univ. Chicago Press, 2005).

    23.
    Watts, D. J. & Strogatz, S. H. Collective dynamics of ‘small-world’ networks. Nature 393, 440–442 (1998).
    ADS  CAS  PubMed  MATH  Google Scholar 

    24.
    Medeiros, L. P., Garcia, G., Thompson, J. N. & Guimarães, P. R. Jr. The geographic mosaic of coevolution in mutualistic networks. Proc. Natl Acad. Sci. USA 115, 12017–12022 (2018).
    CAS  PubMed  Google Scholar 

    25.
    Poisot, T., Canard, E., Mouillot, D., Mouquet, N. & Gravel, D. The dissimilarity of species interaction networks. Ecol. Lett. 15, 1353–1361 (2012).
    PubMed  Google Scholar 

    26.
    Wang, S. & Loreau, M. Biodiversity and ecosystem stability across scales in metacommunities. Ecol. Lett. 19, 510–518 (2016).
    PubMed  PubMed Central  Google Scholar 

    27.
    Simberloff, D. & Von Holle, B. Positive interactions of nonindigenous species: invasional meltdown? Biol. Invasions 1, 21–32 (1999).
    Google Scholar 

    28.
    Tylianakis, J. M., Laliberté, E., Nielsen, A. & Bascompte, J. Conservation of species interaction networks. Biol. Conserv. (2010).

    29.
    Kennedy, C. M., Oakleaf, J. R., Theobald, D. M., Baruch-Mordo, S. & Kiesecker, J. Managing the middle: a shift in conservation priorities based on the global human modification gradient. Glob. Chang. Biol. 25, 811–826 (2019).
    Google Scholar 

    30.
    Bastin, J.-F. et al. The global tree restoration potential. Science 365, 76–79 (2019).
    ADS  CAS  PubMed  Google Scholar 

    31.
    Rezende, E. L., Lavabre, J. E., Guimarães, P. R., Jordano, P. & Bascompte, J. Non-random coextinctions in phylogenetically structured mutualistic networks. Nature 448, 925–928 (2007).
    ADS  CAS  PubMed  Google Scholar 

    32.
    Mello, M. A. R. et al. The missing part of seed dispersal networks: structure and robustness of bat-fruit interactions. PLoS One 6, e17395 (2011).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    33.
    Schleuning, M. et al. Ecological, historical and evolutionary determinants of modularity in weighted seed-dispersal networks. Ecol. Lett. 17, 454–463 (2014).
    PubMed  Google Scholar 

    34.
    Chen, S. C. & Moles, A. T. A mammoth mouthful? A test of the idea that larger animals ingest larger seeds. Glob. Ecol. Biogeogr. 24, 1269–1280 (2015).

    35.
    Dalsgaard, B. et al. Opposed latitudinal patterns of network-derived and dietary specialization in avian plant–frugivore interaction systems. Ecography 40, 1395–1401 (2016).
    Google Scholar 

    36.
    de Assis Bomfim, J., Guimarães, P. R., Peres, C. A., Carvalho, G. & Cazetta, E. Local extinctions of obligate frugivores and patch size reduction disrupt the structure of seed dispersal networks. Ecography 41, 1899–1909 (2018).
    Google Scholar 

    37.
    Castaño, J. H., Carranza, J. A. & Pérez-Torres, J. Diet and trophic structure in assemblages of montane frugivorous phyllostomid bats. Acta Oecol. 91, 81–90 (2018).
    ADS  Google Scholar 

    38.
    Escribano-Avila, G., Lara-Romero, C., Heleno, R. & Traveset, A. in Ecological Networks in the Tropics (eds Dáttilo, W. & Rico-Gray, V.) 93–110 (Springer, 2018).

    39.
    Laurindo, R. S., Novaes, R. L. M., Vizentin-Bugoni, J. & Gregorin, R. The effects of habitat loss on bat-fruit networks. Biodivers. Conserv. 28, 589–601 (2019).
    Google Scholar 

    40.
    Dugger, P. J. et al. Seed-dispersal networks are more specialized in the Neotropics than in the Afrotropics. Glob. Ecol. Biogeogr. 28, 248–261 (2019).
    Google Scholar 

    41.
    Cayuela, L., Granzow-de la Cerda, Í., Albuquerque, F. S. & Golicher, D. J. Taxonstand: an R package for species names standardisation in vegetation databases. Methods Ecol. Evol. 3, 1078–1083 (2012).
    Google Scholar 

    42.
    Boyle, B. et al. The taxonomic name resolution service: an online tool for automated standardization of plant names. BMC Bioinformatics 14, 16 (2013).
    PubMed  PubMed Central  Google Scholar 

    43.
    Trøjelsgaard, K., Heleno, R. & Traveset, A. Native and alien flower visitors differ in partner fidelity and network integration. Ecol. Lett. 22, 1264–1273 (2019).
    PubMed  Google Scholar 

    44.
    Aslan, C. E. Implications of non-native species for mutualistic network resistance and resilience. PLoS One 14, e0217498 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    45.
    Araujo, A. C. et al. Spatial distance and climate determine modularity in a cross-biomes plant–hummingbird interaction network in Brazil. J. Biogeogr. 45, 1846–1858 (2018).
    Google Scholar 

    46.
    Emer, C. et al. Seed-dispersal interactions in fragmented landscapes—a metanetwork approach. Ecol. Lett. 21, 484–493 (2018).
    PubMed  Google Scholar 

    47.
    Nnakenyi, C. A., Traveset, A., Heleno, R., Minoarivelo, H. O. & Hui, C. Fine-tuning the nested structure of pollination networks by adaptive interaction switching, biogeography and sampling effect in the Galápagos Islands. Oikos 128, 1413–1423 (2019).
    Google Scholar 

    48.
    Dunne, J. A. in Ecological Networks: Linking Structure to Dynamics in Food Webs (eds Pascual, M. & Dunne, J. A.) 27–86 (Oxford Univ. Press, 2006).

    49.
    Briatte, F. ggnetwork: geometries to plot networks with ‘ggplot2’. R package v.0.5.1. (R Foundation for Statistical Computing, 2016).

    50.
    Becker, R. A., Wilks, A. R., Brownrigg, R., Minka, T. P. & Deckmyn, A. maps: draw geographical maps. R package v.3 (R Foundation for Statistical Computing, 2013).

    51.
    Clauset, A., Newman, M. E. J. & Moore, C. Finding community structure in very large networks. Phys. Rev. E. 70, 066111 (2004).
    ADS  Google Scholar 

    52.
    Csardi, G. & Nepusz, T. The igraph software package for complex network research. InterJournal 1695 (2006).

    53.
    Dinerstein, E. et al. An ecoregion-based approach to protecting half the terrestrial realm. Bioscience 67, 534–545 (2017).
    PubMed  PubMed Central  Google Scholar 

    54.
    Warton, D. I., Duursma, R. A., Falster, D. S. & Taskinen, S. smatr 3—an R package for estimation and inference about allometric lines. Methods Ecol. Evol. 3, 257–259 (2012).
    Google Scholar 

    55.
    Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).

    56.
    Dormann, C. F., Gruber, B. & Fründ, J. Introducing the bipartite package: analysing ecological networks. Interaction 1, 0.2413793 (2008).
    Google Scholar 

    57.
    Bascompte, J., Jordano, P., Melián, C. J. & Olesen, J. M. The nested assembly of plant-animal mutualistic networks. Proc. Natl Acad. Sci. USA 100, 9383–9387 (2003).
    ADS  CAS  PubMed  Google Scholar 

    58.
    Marquitti, F. M. D., Guimarães, P. R., Pires, M. M. & Bittencourt, L. F. MODULAR: software for the autonomous computation of modularity in large network sets. Ecography 37, 221–224 (2014).
    Google Scholar 

    59.
    Newman, M. E. J. & Girvan, M. Finding and evaluating community structure in networks. Phys. Rev. E 69, 026113 (2004).
    ADS  CAS  Google Scholar 

    60.
    Geraci, M. Linear quantile mixed models: the lqmm package for Laplace quantile regression. J. Stat. Softw. 57, 1–29 (2014). More

  • in

    Subclonal reconstruction of tumors by using machine learning and population genetics

    1.
    Greaves, M. & Maley, C. C. Clonal evolution in cancer. Nature 481, 306–313 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 
    2.
    Turajlic, S., Sottoriva, A., Graham, T. & Swanton, C. Resolving genetic heterogeneity in cancer. Nat. Rev. Genet. 20, 404–416 (2019).
    CAS  PubMed  Google Scholar 

    3.
    Nik-Zainal, S. et al. The life history of 21 breast cancers. Cell 149, 994–1007 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    4.
    Dentro, S. C., Wedge, D. C. & Van Loo, P. Principles of reconstructing the subclonal architecture of cancers. Cold Spring Harb. Perspect. Med. 7, a026625 (2017).
    PubMed  PubMed Central  Google Scholar 

    5.
    Roth, A. et al. PyClone: statistical inference of clonal population structure in cancer. Nat. Meth. 11, 396–398 (2014).
    CAS  Google Scholar 

    6.
    Deshwar, A. G. et al. PhyloWGS: reconstructing subclonal composition and evolution from whole-genome sequencing of tumors. Genome Biol. 16, 35 (2015).
    PubMed  PubMed Central  Google Scholar 

    7.
    Miller, C. A. et al. SciClone: inferring clonal architecture and tracking the spatial and temporal patterns of tumor evolution. PLoS Comput. Biol. 10, e1003665 (2014).
    PubMed  PubMed Central  Google Scholar 

    8.
    Lynch, M. et al. Genetic drift, selection and the evolution of the mutation rate. Nat. Rev. Genet. 17, 704–714 (2016).
    CAS  PubMed  Google Scholar 

    9.
    Williams, M. J., Werner, B., Barnes, C. P., Graham, T. A. & Sottoriva, A. Identification of neutral tumor evolution across cancer types. Nat. Genet. 48, 238–244 (2016).
    CAS  PubMed  PubMed Central  Google Scholar 

    10.
    Kessler, D. A. & Levine, H. Large population solution of the stochastic Luria–Delbruck evolution model. Proc. Natl Acad. Sci. USA 110, 11682–11687 (2013).
    CAS  PubMed  Google Scholar 

    11.
    Kessler, D. A. & Levine, H. Scaling solution in the large population limit of the general asymmetric stochastic Luria–Delbrück evolution process. J. Stat. Phys. 158, 783–805 (2015).
    PubMed  Google Scholar 

    12.
    Durrett, R. Population genetics of neutral mutations in exponentially growing cancer cell populations. Ann. Appl. Probabil. 23, 230–250 (2013).
    Google Scholar 

    13.
    Nicholson, M. D. & Antal, T. Universal asymptotic clone size distribution for general population growth. Bull. Math. Biol. 78, 2243–2276 (2016).
    PubMed  PubMed Central  Google Scholar 

    14.
    Griffiths, R. C. & Tavaré, S. The age of a mutation in a general coalescent. Stoch. Models 14, 273–295 (1998).
    Google Scholar 

    15.
    Sun, R. et al. Between-region genetic divergence reflects the mode and tempo of tumor evolution. Nat. Genet. 49, 1015–1024 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    16.
    Williams, M. J. et al. Quantification of subclonal selection in cancer from bulk sequencing data. Nat. Genet. 50, 895–903 (2018).
    CAS  PubMed  PubMed Central  Google Scholar 

    17.
    Hartl, D. L. & Clark, A. G. Principles of Population Genetics (Sinauer Associates, Inc., 2006).

    18.
    Luria, S. E. & Delbrück, M. Mutations of bacteria from virus sensitivity to virus resistance. Genetics 28, 491–511 (1943).
    CAS  PubMed  PubMed Central  Google Scholar 

    19.
    Graham, T. A. & Sottoriva, A. Measuring cancer evolution from the genome. J. Pathol. 241, 183–191 (2017).
    PubMed  Google Scholar 

    20.
    Griffith, M. et al. Optimizing cancer genome sequencing and analysis. Cell Systems 1, 210–223 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    21.
    Cross, W. et al. The evolutionary landscape of colorectal tumorigenesis. Nat. Ecol. Evol. 2, 1661–1672 (2018).
    PubMed  PubMed Central  Google Scholar 

    22.
    Martincorena, I. et al. Universal patterns of selection in cancer and somatic tissues. Cell 171, 1–13 (2017).
    Google Scholar 

    23.
    Zapata, L. et al. Negative selection in tumor genome evolution acts on essential cellular functions and the immunopeptidome. Genome Biol. 19, 924 (2018).
    Google Scholar 

    24.
    Lee, J. J.-K. et al. Tracing oncogene rearrangements in the mutational history of lung adenocarcinoma. Cell 177, 1842–1857.e21 (2019).
    CAS  PubMed  Google Scholar 

    25.
    The ICGC/TCGA Pan-Cancer Analysis of Whole Genomes Consortium. Pan-cancer analysis of whole genomes. Nature 578, 82–93 (2020).
    CAS  Google Scholar 

    26.
    Gerstung, M. et al. The evolutionary history of 2,658 cancers. Nature 578, 122–128 (2020).
    CAS  PubMed  PubMed Central  Google Scholar 

    27.
    Williams, M. J. et al. Measuring the distribution of fitness effects in somatic evolution by combining clonal dynamics with dN/dS ratios. eLife Sci. 9, 612 (2020).
    Google Scholar 

    28.
    Körber, V. et al. Evolutionary trajectories of IDHWT glioblastomas reveal a common path of early tumorigenesis instigated years ahead of initial diagnosis. Cancer Cell 35, 692–704.e12 (2019).
    PubMed  Google Scholar 

    29.
    Barthel, F. P. et al. Longitudinal molecular trajectories of diffuse glioma in adults. Nature 576, 112–120 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    30.
    Shah, S. P. et al. The clonal and mutational evolution spectrum of primary triple-negative breast cancers. Nature 486, 395–399 (2012).
    CAS  PubMed  Google Scholar 

    31.
    Andor, N. et al. Pan-cancer analysis of the extent and consequences of intratumor heterogeneity. Nat. Med. 22, 105–113 (2016).
    CAS  PubMed  Google Scholar 

    32.
    Morris, L. G. T. et al. Pan-cancer analysis of intratumor heterogeneity as a prognostic determinant of survival. Oncotarget 7, 10051–10063 (2016).
    PubMed  PubMed Central  Google Scholar 

    33.
    Jamal-Hanjani, M. et al. Tracking the evolution of non-small-cell lung cancer. N. Engl. J. Med. 376, 2109–2121 (2017).
    CAS  PubMed  Google Scholar 

    34.
    Espiritu, S. M. G. et al. The evolutionary landscape of localized prostate cancers drives clinical aggression. Cell 173, 1003–1013.e15 (2018).
    CAS  PubMed  Google Scholar 

    35.
    Salcedo, A. et al. A community effort to create standards for evaluating tumor subclonal reconstruction. Nat. Biotechnol. 38, 97–107 (2020).
    CAS  PubMed  PubMed Central  Google Scholar 

    36.
    Yang, L. et al. An enhanced genetic model of colorectal cancer progression history. Genome Biol. 20, 168 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    37.
    Yates, L. R. et al. Genomic evolution of breast cancer metastasis and relapse. Cancer Cell 32, 169–184.e7 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    38.
    Gundem, G. et al. The evolutionary history of lethal metastatic prostate cancer. Nature 520, 353–357 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    39.
    Noorani, A. et al. Genomic evidence supports a clonal diaspora model for metastases of esophageal adenocarcinoma. Nat. Genet. 347, 1–10 (2020).
    Google Scholar 

    40.
    Navin, N. E. The first five years of single-cell cancer genomics and beyond. Genome Res. 25, 1499–1507 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    41.
    Chkhaidze, K. et al. Spatially constrained tumour growth affects the patterns of clonal selection and neutral drift in cancer genomic data. PLoS Comput. Biol. 15, e1007243 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    42.
    Fusco, D., Gralka, M., Kayser, J., Anderson, A. & Hallatschek, O. Excess of mutational jackpot events in expanding populations revealed by spatial Luria–Delbrück experiments. Nat. Commun. 7, 12760 (2016).
    CAS  PubMed  PubMed Central  Google Scholar 

    43.
    Teh, Y. W. Dirichlet processes. in Encyclopedia of Machine Learning (eds Sammut, C. & Webb, G.) 280–287 (Springer, 2011).

    44.
    Ghahramani, Z., Jordan, M. I. & Adams, R. P. Tree-structured stick breaking for hierarchical data. in Advances in Neural Information Processing Systems (eds Lafferty, J. D. et al.) 2319–2327 (Neural Information Processing Systems, 2010).

    45.
    Ma, Z. & Leijon, A. Bayesian estimation of beta mixture models with variational inference. IEEE Trans. Pattern Anal. Mach. Intell. 33, 2160–2173 (2011).
    PubMed  Google Scholar 

    46.
    Clauset, A., Shalizi, C. R. & Newman, M. E. J. Power-law distributions in empirical data. SIAM Rev. 51, 661–703 (2009).
    Google Scholar 

    47.
    Schröder, C. & Rahmann, S. A hybrid parameter estimation algorithm for beta mixtures and applications to methylation state classification. Algorithms Mol. Biol. 12, 21 (2017).
    PubMed  PubMed Central  Google Scholar 

    48.
    Biernacki, C., Celeux, G. & Govaert, G. Assessing a mixture model for clustering with the integrated completed likelihood. IEEE Trans. Pattern Anal. Mach. Intell. 22, 719–725 (2000).
    Google Scholar  More

  • in

    Flow-through stable isotope probing (Flow-SIP) minimizes cross-feeding in complex microbial communities

    1.
    Boschker HTS, Nold SC, Wellsbury P, Bos D, De Graaf W, Pel R, et al. Direct linking of microbial populations to specific biogeochemical processes by 13C-labelling of biomarkers. Nature. 1998;392:801–4.
    CAS  Article  Google Scholar 
    2.
    Radajewski S, Ineson P, Parekh NR, Murrell JC. Stable-isotope probing as a tool in microbial ecology. Nature. 2000;403:646–9.
    CAS  Article  Google Scholar 

    3.
    Orphan VJ, House CH, Hinrichs K-U, McKeegan K, Delong EF. Methane-consuming archaea revealed by directly coupled isotopic and phylogenetic analysis. Science. 2001;293:484–6.
    CAS  Article  Google Scholar 

    4.
    Manefield M, Whiteley AS, Griffiths RI, Bailey MJ. RNA stable isotope probing, a novel means of linking microbial community function to phylogeny. Appl Environ Microbiol. 2002;68:5367–73.
    CAS  Article  Google Scholar 

    5.
    Daebeler A, Bodelier PLE, Yan Z, Hefting MM, Jia Z, Laanbroek HJ. Interactions between Thaumarchaea, Nitrospira and methanotrophs modulate autotrophic nitrification in volcanic grassland soil. ISME J. 2014;8:1–14.
    Article  Google Scholar 

    6.
    Gülay A, Fowler JS, Tatari K, Thamdrup B, Albrechtsen HJ, Abu Al-Soud W, et al. DNA- and RNA-SIP Reveal Nitrospira spp. as Key Drivers of Nitrification in Groundwater-Fed Biofilter. MBio. 2019;10:e01870–19.
    Article  Google Scholar 

    7.
    Berg JS, Pjevac P, Sommer T, Buckner CRT, Philippi M, Hach PF, et al. Dark aerobic sulfide oxidation by anoxygenic phototrophs in anoxic waters. Environ Microbiol. 2019;21:1611–26.
    CAS  Article  Google Scholar 

    8.
    Middelburg JJ, Barranguet C, Boschker HTS, Herman PMJ, Moens T, Heip CHR. The fate of intertidal microphytobenthos carbon: An in situ 13C-labeling study. Limnol Oceanogr. 2000;45:1224–34.
    CAS  Article  Google Scholar 

    9.
    DeRito CM, Pumphrey GM, Madsen EL. Use of field-based stable isotope probing to identify adapted populations and track carbon flow through a phenol-degrading soil microbial community. Appl Environ Microbiol. 2005;71:7858–65.
    CAS  Article  Google Scholar 

    10.
    Dumont MG, Pommerenke B, Casper P, Conrad R. DNA-, rRNA- and mRNA-based stable isotope probing of aerobic methanotrophs in lake sediment. Environ Microbiol. 2011;13:1153–67.
    CAS  Article  Google Scholar 

    11.
    Dolinšek J, Lagkouvardos I, Wanek W, Wagner M, Daims H. Interactions of nitrifying bacteria and heterotrophs: Identification of a Micavibrio-like putative predator of Nitrospira spp. Appl Environ Microbiol. 2013;79:2027–37.
    Article  Google Scholar 

    12.
    Neufeld JD, Schäfer H, Cox MJ, Boden R, McDonald IR, Murrell JC. Stable-isotope probing implicates Methylophaga spp and novel Gammaproteobacteria in marine methanol and methylamine metabolism. ISME J. 2007;1:480–91.
    CAS  Article  Google Scholar 

    13.
    Ho A, Angel R, Veraart AJ, Daebeler A, Jia Z, Kim SY, et al. Biotic interactions in microbial communities as modulators of biogeochemical processes: Methanotrophy as a model system. Front Microbiol. 2016;7:1–11.
    Article  Google Scholar 

    14.
    Lueders T, Manefield M, Friedrich MW. Enhanced sensitivity of DNA- and rRNA-based stable isotope probing by fractionation and quantitative analysis of isopycnic centrifugation gradients. Environ Microbiol. 2004;6:73–78.
    CAS  Article  Google Scholar 

    15.
    Maxfield PJ, Hornibrook ERC, Evershed RP. Estimating high-affinity methanotrophic bacterial biomass, growth, and turnover in soil by phospholipid fatty acid 13C labeling. Appl Environ Microbiol. 2006;72:3901–7.
    CAS  Article  Google Scholar 

    16.
    Pan C, Fischer CR, Hyatt D, Bowen BP, Hettich RL, Banfield JF. Quantitative tracking of isotope flows in proteomes of microbial communities. Mol Cell Proteom. 2011;10:1–11.
    Google Scholar 

    17.
    Albertsen M, Hansen LBS, Saunders AM, Nielsen PH, Nielsen KL. A metagenome of a full-scale microbial community carrying out enhanced biological phosphorus removal. ISME J. 2012;6:1094–106.
    CAS  Article  Google Scholar 

    18.
    Munck C, Albertsen M, Telke A, Ellabaan M, Nielsen PH, Sommer MOA. Limited dissemination of the wastewater treatment plant core resistome. Nat Commun. 2015;6:8452.
    CAS  Article  Google Scholar 

    19.
    Krümmel A, Harms H. Effect of organic matter on growth and cell yield of ammonia-oxidizing bacteria. Arch Microbiol. 1982;133:50–54.
    Article  Google Scholar 

    20.
    Spieck E, Lipski A. Cultivation, growth physiology, and chemotaxonomy of nitrite-oxidizing bacteria. In: Klotz MG, editor. Methods in enzymology, 1st ed. San Diego, USA: Elsevier Inc.; 2011. pp 109–30.

    21.
    Li W. Estimating heterotrophic bacterial productivity by inorganic radiocarbon uptake: importance of establishing time courses of uptake. Mar Ecol Prog Ser. 1982;8:167–72.
    Article  Google Scholar 

    22.
    Roslev P, Larsen MB, Jørgensen D, Hesselsoe M. Use of heterotrophic CO2 assimilation as a measure of metabolic activity in planktonic and sessile bacteria. J Microbiol Methods. 2004;59:381–93.
    CAS  Article  Google Scholar 

    23.
    Okabe S, Kindaichi T, Ito T. Fate of 14C-labeled microbial products derived from nitrifying bacteria in autotrophic nitrifying biofilms. Appl Environ Microbiol. 2005;71:3987–94.
    CAS  Article  Google Scholar 

    24.
    Dirnhuber P, Schütz F. The isomeric transformation of urea into ammonium cyanate in aqueous solutions. Biochem J. 1948;42:628–32.
    CAS  Article  Google Scholar 

    25.
    Palatinszky M, Herbold C, Jehmlich N, Pogoda M, Han P, Von BergenM, et al. Cyanate as an energy source for nitrifiers. Nature. 2015;524:105–8.
    CAS  Article  Google Scholar 

    26.
    Kitzinger K, Padilla CC, Marchant HK, Hach PF, Herbold CW, Kidane AT, et al. Cyanate and urea are substrates for nitrification by Thaumarchaeota in the marine environment. Nat Microbiol. 2019;4:234–43.
    CAS  Article  Google Scholar 

    27.
    Hatzenpichler R, Scheller S, Tavormina PL, Babin BM, Tirrell DA, Orphan VJ. In situ visualization of newly synthesized proteins in environmental microbes using amino acid tagging and click chemistry. Environ Microbiol. 2014;16:2568–90.
    CAS  Article  Google Scholar  More

  • in

    Centers of endemism of freshwater protists deviate from pattern of taxon richness on a continental scale

    1.
    Segawa, T. et al. Bipolar dispersal of red-snow algae. Nat. Commun. 9, 3094. https://doi.org/10.1038/s41467-018-05521-w (2018).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 
    2.
    Tedersoo, L. et al. Fungal biogeography. Global diversity and geography of soil fungi. Science (New York, N.Y.) 346, 1256688. https://doi.org/10.1126/science.1256688 (2014).
    CAS  Article  Google Scholar 

    3.
    Dunthorn, M., Stoeck, T., Wolf, K., Breiner, H.-W. & Foissner, W. Diversity and endemism of ciliates inhabiting Neotropical phytotelmata. Syst. Biodivers. 10, 195–205. https://doi.org/10.1080/14772000.2012.685195 (2012).
    Article  Google Scholar 

    4.
    Siver, P. A. & Lott, A. M. Biogeographic patterns in scaled chrysophytes from the east coast of North. America 57, 451–466. https://doi.org/10.1111/j.1365-2427.2011.02711.x (2012).
    Article  Google Scholar 

    5.
    Gaston, K. J. Global patterns in biodiversity. Nature 405, 220–227. https://doi.org/10.1038/35012228 (2000).
    CAS  Article  PubMed  Google Scholar 

    6.
    Bass, D., Boenigk, J. & Fontaneto, D. In Biogeography of Microscopic Organisms (ed. Fontaneto, D.) 88–110 (Cambridge University Press, Cambridge, 2011).
    Google Scholar 

    7.
    Caron, D. A. Past President’s address: protistan biogeography: why all the fuss?. J. Eukaryot. Microbiol. 56, 105–112. https://doi.org/10.1111/j.1550-7408.2008.00381.x (2009).
    ADS  Article  PubMed  Google Scholar 

    8.
    Foissner, W. Biogeography and dispersal of micro-organisms: a review emphasizing protists (2006).

    9.
    Coesel, P. F. M. & Krienitz, L. Diversity and geographic distribution of desmids and other coccoid green algae. Biodivers. Conserv. 17, 381–392. https://doi.org/10.1007/s10531-007-9256-5 (2008).
    Article  Google Scholar 

    10.
    Darling, K. F. & Wade, C. M. The genetic diversity of planktic foraminifera and the global distribution of ribosomal RNA genotypes. Mar. Micropaleontol. 67, 216–238. https://doi.org/10.1016/j.marmicro.2008.01.009 (2008).
    ADS  Article  Google Scholar 

    11.
    Vanormelingen, P., Verleyen, E. & Vyverman, W. The diversity and distribution of diatoms: from cosmopolitanism to narrow endemism. Biodivers. Conserv. 17, 393–405. https://doi.org/10.1007/s10531-007-9257-4 (2008).
    Article  Google Scholar 

    12.
    Stoeck, T., Bruemmer, F. & Foissner, W. Evidence for local ciliate endemism in an alpine anoxic lake. Microbiol. Ecol. 54, 478–486. https://doi.org/10.1007/s00248-007-9213-6 (2007).
    Article  Google Scholar 

    13.
    Fernández, L. D., Hernández, C. E., Schiaffino, M. R., Izaguirre, I. & Lara, E. Geographical distance and local environmental conditions drive the genetic population structure of a freshwater microalga (Bathycoccaceae; Chlorophyta) in Patagonian lakes. FEMS Microbiol. Ecol. 93, 37. https://doi.org/10.1093/femsec/fix125 (2017).
    CAS  Article  Google Scholar 

    14.
    Filker, S., Sommaruga, R., Vila, I. & Stoeck, T. Microbial eukaryote plankton communities of high-mountain lakes from three continents exhibit strong biogeographic patterns. Mol. Ecol. 25, 2286–2301. https://doi.org/10.1111/mec.13633 (2016).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    15.
    de Vargas, C. et al. Eukaryotic plankton diversity in the sunlit ocean. Science 348, 1261605. https://doi.org/10.1126/science.1261605 (2015).
    CAS  Article  PubMed  Google Scholar 

    16.
    Ibarbalz, F. M. et al. Global trends in marine plankton diversity across kingdoms of life. Cell 179, 1084-1097.e21. https://doi.org/10.1016/j.cell.2019.10.008 (2019).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    17.
    Bock, C., Salcher, M., Jensen, M., Pandey, R. V. & Boenigk, J. Synchrony of eukaryotic and prokaryotic planktonic communities in three seasonally sampled Austrian lakes. Front. Microbiol. 9, 1290. https://doi.org/10.3389/fmicb.2018.01290 (2018).
    Article  PubMed  PubMed Central  Google Scholar 

    18.
    Boenigk, J. et al. Geographic distance and mountain ranges structure freshwater protist communities on a European scale. Metabarcoding and Metagenomics 2, e21519. https://doi.org/10.3897/mbmg.2.21519 (2018).
    Article  Google Scholar 

    19.
    He, F. et al. Elevation, aspect, and local environment jointly determine diatom and macroinvertebrate diversity in the Cangshan Mountain, Southwest China. Ecol. Indic. 108, 105618. https://doi.org/10.1016/j.ecolind.2019.105618 (2020).
    Article  Google Scholar 

    20.
    Shen, C. et al. Contrasting elevational diversity patterns between eukaryotic soil microbes and plants. Ecology 95, 3190–3202. https://doi.org/10.1890/14-0310.1 (2014).
    Article  Google Scholar 

    21.
    Bryant, J. A. et al. Colloquium paper: microbes on mountainsides: contrasting elevational patterns of bacterial and plant diversity. Proc. Natl. Acad. Sci. USA 105(Suppl 1), 11505–11511. https://doi.org/10.1073/pnas.0801920105 (2008).
    ADS  Article  PubMed  Google Scholar 

    22.
    McCain, C. M. Could temperature and water availability drive elevational species richness patterns? A global case study for bats. Global Ecol. Biogeogr. 16, 1–13. https://doi.org/10.1111/j.1466-8238.2006.00263.x (2007).
    Article  Google Scholar 

    23.
    Desmond, A. Janet Browne, The secular ark. Studies in the history of biogeography, New Haven, Conn., and London, Yale University Press, 1983, 8vo, pp. x, 273, illus., £21.00. Med. Hist. 27, 452–453. https://doi.org/10.1017/s0025727300043611 (1983).
    Article  PubMed Central  Google Scholar 

    24.
    Nemcová, Y., Kreidlová, J., Kosová, A. & Neustupa, J. Lakes and pools of Aquitaine region (France)—a biodiversity hotspot of Synurales in Europe. Nova Hedw 95, 1–24. https://doi.org/10.1127/0029-5035/2012/0036 (2012).
    Article  Google Scholar 

    25.
    van de Vijver, B., Gremmen, N. J. M. & Beyens, L. The genus Stauroneis (Bacillariophyceae) in the Antarctic region. J. Biogeogr. 32, 1791–1798. https://doi.org/10.1111/j.1365-2699.2005.01325.x (2005).
    Article  Google Scholar 

    26.
    Martiny, J. B. H. et al. Microbial biogeography: putting microorganisms on the map. Nat. Rev. Microbiol. 4, 102–112. https://doi.org/10.1038/nrmicro1341 (2006).
    CAS  Article  PubMed  Google Scholar 

    27.
    Lepère, C. et al. Geographic distance and ecosystem size determine the distribution of smallest protists in lacustrine ecosystems. FEMS Microbiol. Ecol. 85, 85–94. https://doi.org/10.1111/1574-6941.12100 (2013).
    Article  PubMed  Google Scholar 

    28.
    Green, J. L. et al. Spatial scaling of microbial eukaryote diversity. Nature 432, 747–750. https://doi.org/10.1038/nature03034 (2004).
    ADS  CAS  Article  PubMed  Google Scholar 

    29.
    Lara, E., Roussel-Delif, L., Fournier, B., Wilkinson, D. M. & Mitchell, E. A. D. Soil microorganisms behave like macroscopic organisms: patterns in the global distribution of soil euglyphid testate amoebae. J. Biogeogr. 43, 520–532. https://doi.org/10.1111/jbi.12660 (2016).
    Article  Google Scholar 

    30.
    Kier, G. et al. A global assessment of endemism and species richness across island and mainland regions. Proc. Natl. Acad. Sci. USA 106, 9322–9327. https://doi.org/10.1073/pnas.0810306106 (2009).
    ADS  Article  PubMed  Google Scholar 

    31.
    Orme, C. D. L. et al. Global hotspots of species richness are not congruent with endemism or threat. Nature 436, 1016–1019. https://doi.org/10.1038/nature03850 (2005).
    ADS  CAS  Article  PubMed  Google Scholar 

    32.
    Schmitt, T. Biogeographical and evolutionary importance of the European high mountain systems. Front. Zool. 6, 9. https://doi.org/10.1186/1742-9994-6-9 (2009).
    Article  PubMed  PubMed Central  Google Scholar 

    33.
    Vimercati, L., Darcy, J. L. & Schmidt, S. K. The disappearing periglacial ecosystem atop Mt. Kilimanjaro supports both cosmopolitan and endemic microbial communities. Sci. Rep. 9, 10676. https://doi.org/10.1038/s41598-019-46521-0 (2019).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    34.
    McCain, C. M. Elevational gradients in diversity of small mammals. Ecology 86, 366–372. https://doi.org/10.1890/03-3147 (2005).
    Article  Google Scholar 

    35.
    Malviya, S. et al. Insights into global diatom distribution and diversity in the world’s ocean. Proc. Natl. Acad. Sci. USA 113, E1516–E1525. https://doi.org/10.1073/pnas.1509523113 (2016).
    CAS  Article  PubMed  Google Scholar 

    36.
    Khomich, M., Kauserud, H., Logares, R., Rasconi, S. & Andersen, T. Planktonic protistan communities in lakes along a large-scale environmental gradient. FEMS Microbiol. Ecol. https://doi.org/10.1093/femsec/fiw231 (2017).
    Article  PubMed  Google Scholar 

    37.
    Škaloud, P. et al. Speciation in protists: Spatial and ecological divergence processes cause rapid species diversification in a freshwater chrysophyte. Mol. Ecol. 28, 1084–1095. https://doi.org/10.1111/mec.15011 (2019).
    Article  PubMed  Google Scholar 

    38.
    Godhe, A., McQuoid, M. R., Karunasagar, I., Karunasagar, I. & Rehnstam-Holm, A.-S. Comparison of three common molecular tools for distinguishing among geographically separated clones of the diatom Skeletonema marinoi sarno et zingone (Bacillariophyceae). J. Phycol. 42, 280–291. https://doi.org/10.1111/j.1529-8817.2006.00197.x (2006).
    CAS  Article  Google Scholar 

    39.
    Jobst, J., King, K. & Hemleben, V. Molecular evolution of the internal transcribed spacers (ITS1 and ITS2) and phylogenetic relationships among species of the family cucurbitaceae. Mol. Phylogenet. Evol. 9, 204–219. https://doi.org/10.1006/mpev.1997.0465 (1998).
    CAS  Article  PubMed  Google Scholar 

    40.
    Lobo-Hajdu, G. Intragenomic, intra- and interspecific variation in the rDNA its of Porifera revealed by PCR-singlestrand conformation polymorphism (PCR-SSCP). Bollettino dei Musei e degli Istituti Biologici 68, 413–423 (2004).
    Google Scholar 

    41.
    Needham, D. M., Sachdeva, R. & Fuhrman, J. A. Ecological dynamics and co-occurrence among marine phytoplankton, bacteria and myoviruses shows microdiversity matters. ISME J. 11, 1614–1629. https://doi.org/10.1038/ismej.2017.29 (2017).
    Article  PubMed  PubMed Central  Google Scholar 

    42.
    Derot, J. et al. Response of phytoplankton traits to environmental variables in French lakes: new perspectives for bioindication. Ecol. Indic. 108, 105659. https://doi.org/10.1016/j.ecolind.2019.105659 (2020).
    CAS  Article  Google Scholar 

    43.
    Boo, S. M. et al. Complex phylogeographic patterns in the freshwater alga Synura provide new insights into ubiquity vs. endemism in microbial eukaryotes. Mol. Ecol. 19, 4328–4338. https://doi.org/10.1111/j.1365-294X.2010.04813.x (2010).
    Article  PubMed  Google Scholar 

    44.
    Foissner, W. & Hawksworth, D. L. (eds) Protist Diversity and Geographical Distribution (Springer, Dordrecht, 2009).
    Google Scholar 

    45.
    Schiaffino, M. R. et al. Microbial eukaryote communities exhibit robust biogeographical patterns along a gradient of Patagonian and Antarctic lakes. Environ. Microbiol. 18, 5249–5264. https://doi.org/10.1111/1462-2920.13566 (2016).
    CAS  Article  PubMed  Google Scholar 

    46.
    Boenigk, J. et al. Evidence for geographic isolation and signs of endemism within a protistan morphospecies. Appl. Environ. Microbiol. 72, 5159–5164. https://doi.org/10.1128/AEM.00601-06 (2006).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    47.
    Foissner, W., Chao, A. & Katz, L. A. Diversity and geographic distribution of ciliates (Protista: Ciliophora). Biodivers. Conserv. 17, 345–363. https://doi.org/10.1007/s10531-007-9254-7 (2008).
    Article  Google Scholar 

    48.
    Payo, D. A. et al. Extensive cryptic species diversity and fine-scale endemism in the marine red alga Portieria in the Philippines. Proc. R. Soc. B 280, 20122660. https://doi.org/10.1098/rspb.2012.2660 (2013).
    Article  PubMed  Google Scholar 

    49.
    Siver, P. A., Skogstad, A. & Nemcová, Y. Endemism, palaeoendemism and migration: the case for the ‘European endemic’, Mallomonas intermedia. Eur. J. Phycol. 54, 222–234. https://doi.org/10.1080/09670262.2018.1544377 (2019).
    Article  Google Scholar 

    50.
    Cox, F., Newsham, K. K. & Robinson, C. H. Endemic and cosmopolitan fungal taxa exhibit differential abundances in total and active communities of Antarctic soils. Environ. Microbiol. 21, 1586–1596. https://doi.org/10.1111/1462-2920.14533 (2019).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    51.
    Ibelings, B. W. et al. Host parasite interactions between freshwater phytoplankton and chytrid fungi (Chytridiomycota). J. Phycol. 40, 437–453. https://doi.org/10.1111/j.1529-8817.2004.03117.x (2004).
    Article  Google Scholar 

    52.
    Logares, R. et al. Infrequent marine-freshwater transitions in the microbial world. Trends Microbiol. 17, 414–422. https://doi.org/10.1016/j.tim.2009.05.010 (2009).
    CAS  Article  PubMed  Google Scholar 

    53.
    Hewitt, G. M. The structure of biodiversity—insights from molecular phylogeography. Front. Zool. 1, 4. https://doi.org/10.1186/1742-9994-1-4 (2004).
    Article  PubMed  PubMed Central  Google Scholar 

    54.
    Vetaas, O. R. & Grytnes, J.-A. Distribution of vascular plant species richness and endemic richness along the Himalayan elevation gradient in Nepal. Global Ecol. Biogeogr. 11, 291–301. https://doi.org/10.1046/j.1466-822X.2002.00297.x (2002).
    Article  Google Scholar 

    55.
    Nogués-Bravo, D., Araújo, M. B., Romdal, T. & Rahbek, C. Scale effects and human impact on the elevational species richness gradients. Nature 453, 216–219. https://doi.org/10.1038/nature06812 (2008).
    ADS  CAS  Article  PubMed  Google Scholar 

    56.
    Catalán, J. et al. High mountain lakes: extreme habitats and witnesses of environmental changes. Limnetica 25, 551–584 (2006).
    Google Scholar 

    57.
    Sommaruga, R. The role of solar UV radiation in the ecology of alpine lakes. J. Photochem. Photobiol. B Biol. 62, 35–42. https://doi.org/10.1016/S1011-1344(01)00154-3 (2001).
    CAS  Article  Google Scholar 

    58.
    Morris, D. P. et al. The attenuation of solar UV radiation in lakes and the role of dissolved organic carbon. Limnol. Oceanogr. 40, 1381–1391. https://doi.org/10.4319/lo.1995.40.8.1381 (1995).
    ADS  CAS  Article  Google Scholar 

    59.
    Sommaruga, R. & Augustin, G. Seasonality in UV transparency of an alpine lake is associated to changes in phytoplankton biomass. Aquat. Sci. 68, 129–141. https://doi.org/10.1007/s00027-006-0836-3 (2006).
    Article  Google Scholar 

    60.
    Catalan, J. et al. High mountain lakes: extreme habitats and witnesses of environmental changes. Limnética 25, 551–584 (2006).
    Google Scholar 

    61.
    Ortiz-Álvarez, R., Triadó-Margarit, X., Camarero, L., Casamayor, E. O. & Catalan, J. High planktonic diversity in mountain lakes contains similar contributions of autotrophic, heterotrophic and parasitic eukaryotic life forms. Sci. Rep. 8, 302. https://doi.org/10.1038/s41598-018-22835-3 (2018).
    CAS  Article  Google Scholar 

    62.
    Kammerlander, B. et al. High diversity of protistan plankton communities in remote high mountain lakes in the European Alps and the Himalayan mountains. FEMS Microbiol. Ecol. 91, 429. https://doi.org/10.1093/femsec/fiv010 (2015).
    CAS  Article  Google Scholar 

    63.
    Tartarotti, B. et al. UV-induced DNA damage in Cyclops abyssorum tatricus populations from clear and turbid alpine lakes. J. Plankton Res. 36, 557–566. https://doi.org/10.1093/plankt/fbt109 (2014).
    CAS  Article  PubMed  Google Scholar 

    64.
    Brettum, P. & Halvorsen, G. The phytoplankton of Lake Atnsjøen, Norway—a long-term investigation. Hydrobiologia 521, 141–147. https://doi.org/10.1023/B:HYDR.0000026356.09421.e3 (2004).
    Article  Google Scholar 

    65.
    Karlsson, J. et al. Light limitation of nutrient-poor lake ecosystems. Nature 460, 506–509. https://doi.org/10.1038/nature08179 (2009).
    ADS  CAS  Article  PubMed  Google Scholar 

    66.
    Bergström, A.-K., Karlsson, D., Karlsson, J. & Vrede, T. N-limited consumer growth and low nutrient regeneration N: P ratios in lakes with low N deposition. Ecosphere 6, 9. https://doi.org/10.1890/ES14-00333.1 (2015).
    Article  Google Scholar 

    67.
    Kritzberg, E. S. et al. Browning of freshwaters: consequences to ecosystem services, underlying drivers, and potential mitigation measures. Ambio 49, 375–390. https://doi.org/10.1007/s13280-019-01227-5 (2020).
    Article  PubMed  Google Scholar 

    68.
    Gustafsson, B. G. & Westman, P. On the causes for salinity variations in the Baltic Sea during the last 8500 years. Paleoceanography 17, 12-1-12–14. https://doi.org/10.1029/2000PA000572 (2002).
    Article  Google Scholar 

    69.
    Filker, S., Kühner, S., Heckwolf, M., Dierking, J. & Stoeck, T. A fundamental difference between macrobiota and microbial eukaryotes: protistan plankton has a species maximum in the freshwater-marine transition zone of the Baltic Sea. Environ. Microbiol. 21, 603–617. https://doi.org/10.1111/1462-2920.14502 (2019).
    CAS  Article  PubMed  Google Scholar 

    70.
    Schiewer, U. In Ecology of Baltic Coastal Waters (ed. Schiewer, U.) 395–417 (Springer, Berlin, 2008).
    Google Scholar 

    71.
    Falkowski, P. G. et al. The evolution of modern eukaryotic phytoplankton. Science (New York, N.Y.) 305, 354–360. https://doi.org/10.1126/science.1095964 (2004).
    ADS  CAS  Article  Google Scholar 

    72.
    Cermeño, P., Falkowski, P. G., Romero, O. E., Schaller, M. F. & Vallina, S. M. Continental erosion and the Cenozoic rise of marine diatoms. Proc. Natl. Acad. Sci. USA 112, 4239–4244. https://doi.org/10.1073/pnas.1412883112 (2015).
    ADS  CAS  Article  PubMed  Google Scholar 

    73.
    Rothschild, L. J. The influence of UV radiation on protistan evolution. J. Eukaryot. Microbiol. https://doi.org/10.1111/j.1550-7408.1999.tb06074.x| (1999).
    Article  PubMed  Google Scholar 

    74.
    Rose, J. M. & Caron, D. A. Does low temperature constrain the growth rates of heterotrophic protists? Evidence and implications for algal blooms in cold waters. Limnol. Oceanogr. 52, 886–895. https://doi.org/10.4319/lo.2007.52.2.0886 (2007).
    ADS  Article  Google Scholar 

    75.
    Ægisdóttir, H. H., Kuss, P. & Stöcklin, J. Isolated populations of a rare alpine plant show high genetic diversity and considerable population differentiation. Ann. Bot. 104, 1313–1322. https://doi.org/10.1093/aob/mcp242 (2009).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    76.
    Cain, M. L., Milligan, B. G. & Strand, A. E. Long-distance seed dispersal in plant populations. Am. J. Bot. 87, 1217–1227. https://doi.org/10.2307/2656714 (2000).
    CAS  Article  PubMed  Google Scholar 

    77.
    Nemcová, Y. & Pichrtova, M. Shape dynamics of silica scales (Chrysophyceae, Stramenopiles) associated with pH. Fottea 12, 281–291. https://doi.org/10.5507/fot.2012.020 (2012).
    Article  Google Scholar 

    78.
    Leadbeater, B. S. C. & Green, J. C. Flagellates: Unity, Diversity and Evolution. Chapter 12: Functional Diversity of Heterotrophic Flagellates in Aquatic Ecosystems (CRC Press, Cambridge, 2000).
    Google Scholar 

    79.
    Lange, A. et al. AmpliconDuo: a split-sample filtering protocol for high-throughput amplicon sequencing of microbial communities. PLoS ONE 10, e0141590. https://doi.org/10.1371/journal.pone.0141590 (2015).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    80.
    Fu, L., Niu, B., Zhu, Z., Wu, S. & Li, W. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinformatics 28, 3150–3152. https://doi.org/10.1093/bioinformatics/bts565 (2012).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    81.
    Jensen, M. V9_Clust.R. R-Scrift for modifying DNA-sequence-abundance-tables: clustering of related sequences (e.g. SSU-ITS1) according to 100% identical subsequences. https://github.com/manfred-uni-essen/V9-cluster (2017).

    82.
    Mahé, F., Rognes, T., Quince, C., de Vargas, C. & Dunthorn, M. Swarm v2: highly-scalable and high-resolution amplicon clustering. PeerJ 3, e1420. https://doi.org/10.7717/peerj.1420 (2015).
    Article  PubMed  PubMed Central  Google Scholar 

    83.
    R Core Team. R: A language and environment for statistical computing (2019).

    84.
    Hijmans, R. J. Spherical Trigonometry [R package geosphere version 1.5–10] (2019).

    85.
    Kruskal, W. H. & Wallis, W. A. Use of ranks in one-criterion variance analysis. J. Am. Stat. Assoc. 47, 583–621. https://doi.org/10.1080/01621459.1952.10483441 (1952).
    Article  MATH  Google Scholar 

    86.
    Dunn, O. J. Multiple comparisons among means. J. Am. Stat. Assoc. 56, 52–64. https://doi.org/10.1080/01621459.1961.10482090 (1961).
    MathSciNet  Article  MATH  Google Scholar  More

  • in

    Rare and common vertebrates span a wide spectrum of population trends

    Workflow
    All data syntheses, visualisation and statistical analyses were conducted using R version 3.5.171. For conceptual diagrams of our workflow, see Supplementary Figs. 1 and 2. Effect sizes plotted on graphs were standardised by dividing the effect size by the standard deviation of the corresponding input data.
    Population data
    To quantify vertebrate population change (trends and fluctuations), we extracted the abundance data for 9286 population time series from 2084 species from the publicly available Living Planet Database72 (http://www.livingplanetindex.org/data_portal) that covered the period between 1970 and 2014 (Supplementary Table 1). These time series represent repeated monitoring surveys of the number of individuals in a given area, hereafter, called ‘populations’. Monitoring duration differed among populations, with a mean duration of 23.9 years and a mean sampling frequency of 23.3 time points (Supplementary Fig. 3, see Supplementary Figs. 6 and 7 for effects of monitoring duration on detected trends). In the Living Planet database, 17.9% of populations were sampled annually or in rare cases multiple times per year. The time series we analysed include vertebrate species that span a large variation in age, generation times and other demographic-rate processes. For example, from other work that we have conducted, we have found that when generation time data were available (~50.0% or 484 out of 968 bird species, and 15.6% or 48 out of 306 mammal species), the mean bird generation time is 5.0 years (min = 3.4 years, max = 14.3 years) and the mean mammal generation time is 8.3 years (min = 0.3 years, max = 25 years)45. Thus, we believe that most vertebrate time series within the LPD capture multiple generations.
    In our analysis, we omitted populations which had less than five time points of monitoring data, as previous studies of similar population time series to the ones we analysed have found that shorter time series might not capture directional trends in abundance63. Populations were monitored using different metrics of abundance (e.g., population indices vs number of individuals). Before analysis, we scaled the abundance of each population to a common magnitude between zero and one to analyse within-population relationships to prevent conflating within-population relationships and between-population relationships73. Scaling the abundance data allowed us to explore trends among populations relative to the variation experienced across each time series.
    Phylogenetic data
    We obtained phylogenies for amphibian species from https://vertlife.org4, for bird species from https://birdtree.org8, and for reptile species from https://vertlife.org6. For each of the three classes (Amphibia, Aves and Reptilia), we downloaded 100 trees and randomly chose 10 for analysis (30 trees in total). Species-level phylogenies for the classes Actinopterygii and Mammalia have not yet been resolved with high confidence74,75.
    Rarity metrics, IUCN Red List categories and threat types
    We defined rarity following a simplified version of the ‘seven forms of rarity’ model76, and thus consider rarity to be the state in which species exist when they have a small geographic range, low population size, or narrow habitat specificity. We combined publicly available data from three sources: (1) population records for vertebrate species from the Living Planet Database to calculate mean population size, (2) occurrence data from the Global Biodiversity Information Facility77 (https://www.gbif.org) and range data from BirdLife78 (http://datazone.birdlife.org) to estimate geographic range size and (3) habitat specificity and Red List Category data for each species from the International Union for Conservation79 (https://www.iucnredlist.org). The populations in the Living Planet Database72 do not include species that have gone extinct on a global scale. We extracted the number and types of threats that each species is exposed to globally from their respective species’ IUCN Red List profiles79.
    Quantifying population trends and fluctuations over time
    In the first stage of our analysis, we used state-space models that model abundance (scaled to a common magnitude between zero and one) over time to calculate the amount of overall abundance change experienced by each population (μ,40,80). State-space models account for process noise (σ2) and observation error (τ2) and thus deliver robust estimates of population change when working with large data sets where records were collected using different approaches, such as the Living Planet Database41,81,82. Previous studies have found that not accounting for process noise and measurement error could lead to over-estimation of population declines83, but in our analyses, we found that population trends derived from state-space models were similar to those derived from linear models. Positive μ values indicate population increase and negative μ values indicate population decline. State-space models partition the variance in abundance estimates into estimated process noise (σ2) and observation or measurement error (τ2) and population trends (μ):

    $$X_t = X_{t-1} + mu + varepsilon _t,$$
    (1)

    where Xt and Xt−1 are the scaled (observed) abundance estimates (between 0 and 1) in the present and past year, with process noise represented by εt~ gaussian(0, σ2). We included measurement error following:

    $$Y_t = X_t + F_t,$$
    (2)

    where Yt is the estimate of the true (unobserved) population abundance with measurement error:

    $$F_tsim gaussianleft( {0,,{it{T}}^2} right)$$
    (3)

    We substituted the estimate of population abundance (Yt) into Eq. 1:

    $$Y_{t} = {it{X}}_{{it{t}} – 1} + mu + varepsilon _{it{t}} + {it{F}}_{it{t}}.$$
    (4)

    Given

    $${it{X}}_{{it{t}} – 1} = {it{Y}}_{{it{t}} – 1} – {it{F}}_{{it{t}} – 1}$$
    (5)

    then:

    $${it{Y}}_{it{t}} = {it{Y}}_{t – 1} + mu + varepsilon _t + F_t – F_{t – 1}$$
    (6)

    For comparisons of different approaches to modelling population change, see ‘Comparison of modelling approaches section’.
    Quantifying rarity metrics
    We tested how population change varied across rarity metrics—geographic range, mean population size and habitat specificity – on two different but complementary scales. In the main text, we presented the results of our global-scale analyses, whereas in the SI, we included the results when using only populations from the UK—a country with high monitoring intensity, Thus, we quantified rarity metrics for species monitoring globally and in the UK. The three rarity metrics used in this study were weakly correlated at both UK and global scales (Supplementary Fig. 11).
    Geographic range
    To estimate geographic range for bird species monitored globally, we extracted the area of occurrence in km2 for all bird species in the Living Planet Database that had records in the BirdLife Data Zone78. For mammal species’ geographic range, we used the PanTHERIA database84 (http://esapubs.org/archive/ecol/E090/184/). To estimate geographic range for bony fish, birds, amphibians, mammals and reptiles monitored in the UK (see Supplementary Table 5 for species list), we calculated a km2 occurrence area based on species occurrence data from GBIF77. Extracting and filtering GBIF data and calculating range was computationally intensive and occurrence data availability was lower for certain species. Thus, we did not estimate geographic range from GBIF data for all species part of the Living Planet Database. Instead, we focused on analysing range effects for birds and mammals globally, as they are a very well-studied taxon and for species monitored in the UK, a country with intensive and detailed biodiversity monitoring of vertebrate species. We did not use IUCN range maps, as they were not available for all of our study species, and previous studies using GBIF occurrences to estimate range have found a positive correlation between GBIF-derived and IUCN-derived geographic ranges85.
    For the geographic ranges of species monitored in the UK, we calculated range extent using a minimal convex hull approach based on GBIF occurrence data77. We filtered the GBIF data to remove invalid records and outliers using the CoordinateCleaner package86. We excluded records with no decimal places in the decimal latitude or longitude values, with equal latitude or longitude, within a one-degree radius of the GBIF headquarters in Copenhagen, within 0.0001 degrees of various biodiversity institutions and within 0.1 degrees of capital cities. For each species, we excluded the lower 0.02 and upper 0.98 quantile intervals of the latitude and longitude records to account for outlier points that are records from zoos or other non-wild populations. We drew a convex hull to most parsimoniously encompass all remaining occurrence records using the chull function, and we calculated the area of the resulting polygon using areaPolygon from the geosphere package87.
    Mean size of monitored populations
    We calculated mean size of the monitored population, referred to as population size, across the monitoring duration using the raw abundance data, and we excluded populations, which were not monitored using population counts (i.e., we excluded indexes).
    Habitat specificity
    To create an index of habitat specificity, we extracted the number of distinct habitats a species occupies based on the IUCN habitat category for each species’ profile, accessed through the package rredlist88. We also quantified habitat specificity by surveying the number of breeding and non-breeding habitats for each species from its online IUCN species profile (the ‘habitat and ecology’ section). The two approaches yielded similar results (Supplementary Fig. 10, Supplementary Table 2, key for the profiling method is presented in Supplementary Table 6). We obtained global IUCN Red List Categories and threat types for all study species through their IUCN Red List profiles79.
    Testing the sources of variation in population trends and fluctuations
    In the second stage of our analyses, we modelled the trend and fluctuation estimates from the first stage analyses across latitude, realm, biome, taxa, rarity metrics, phylogenetic relatedness, species’ IUCN Red List Categories and threat type using a Bayesian modelling framework through the package MCMCglmm89. We included a species random intercept effect in the Bayesian models to account for the possible correlation between the trends of populations from the same species (see Supplementary Table 1 for sample sizes). The models ran for 120,000 iterations with a thinning factor of ten, a burn-in period of 20,000 iterations and the default one chain. We assessed model convergence by visually examining trace plots. We used weakly informative priors for all coefficients (an inverse Wishart prior for the variances and a normal prior for the fixed effects):

    $$Prleft( mu right) sim Nleft( {0,,10^8} right)$$
    (7)

    $$Pr(sigma ^2) sim Inverse,Wishart,left( {V = 0,,nu = 0} right)$$
    (8)

    Population trends and fluctuations across latitude, biomes, realms and taxa
    To investigate the geographic and taxonomic patterns of population trends and fluctuations, we modelled population trends (μ) and population fluctuations (σ2), derived from the first stage of our analyses (state-space models), as a function of (1) latitude, (2) realm (freshwater, marine, terrestrial), (3) biome (as defined by the ‘biome’ category in the Living Planet Database, e.g., ‘temperate broadleaf forest’90 and (4) taxa (Actinopterygii, bony fish; Elasmobranchii, sharks and rays; Amphibia, amphibians; Aves, birds; Mammalia, mammals; Reptilia, reptiles). We used separate models for each variable, resulting in four models testing the sources of variation in trends and four additional models focusing on fluctuations. Each categorical model from this second stage of our analyses was fitted with a zero intercept to allow us to determine whether net population trends differed from zero for each of the categories under investigation. The model structures for all models with a categorical fixed effect were identical with the exception of the identity of the fixed effect, and below we describe the taxa model:

    $$mu _{i,j,k} = beta _0 + beta _{0,j} + beta _1 ast taxa_{i,j,k},$$
    (9)

    $$y_{i,j,k}sim gaussianleft( {mu _{i,j,k},sigma ^2} right),$$
    (10)

    where taxai,j,k is the taxa of the ith time series from the jth species; β0 and β1 are the global intercept (in categorical models, β0 = 1) and the slope estimate for the categorical taxa effect (fixed effect), β0j is the species-level departure from β0 (species-level random effect); yi,j,k is the estimate for change in population abundance for the ith population time series from the jth species from the kth taxa.
    Population trends and fluctuations across amphibian, bird and reptile phylogenies
    To determine whether there is a phylogenetic signal in the patterning of population change within amphibian, bird and reptile taxa, we modelled population trends (μ) and fluctuations (σ2) across phylogenetic and species-level taxonomic relatedness. We conducted one model per taxa per population change variable—trends or fluctuations using Bayesian linear mixed effects models using the package MCMCglmm89. We included phylogeny and taxa as random effects. The models did not include fixed effects. We assessed the magnitude of the random effects (phylogeny and species) by inspecting their posterior distributions, with a distribution pushed up against zero indicating lack of effect, as these distributions are always bounded by zero and have only positive values. We used parameter-expanded priors, with a variance-covariance structure that allows the slopes of population trend (the μ values from the first stage analysis using state-space models) to covary for each random effect. The prior and model structure were as follows:

    $$Prleft( mu right) sim Nleft( {0,,10^8} right),$$
    (11)

    $$Prleft( {sigma ^2} right) sim Inverse,Wishart,left( {V = 1,,nu = 1} right),$$
    (12)

    $$mu _{i,k,m} = beta _0 + beta _{0,k} + beta _{0,m},$$
    (13)

    $$y_{i,k,m} sim gaussianleft( {mu _{i,k,m},,sigma ^2} right),$$
    (14)

    where β0 is the global intercept (β0 = 1), β0l is the phylogeny-level departure from β0 (phylogeny random effect); yi,k,m is the estimate for change in population abundance for the ith population time series for the kth species with the mth phylogenetic distance.
    To account for phylogenetic uncertainty for each class, we ran ten models with identical structures, but based on different randomly selected phylogenetic trees. We report the mean estimate and its range for each class.
    Population trends and fluctuations across rarity metrics
    To test the influence of rarity metrics (geographic range, mean population size and habitat specificity) on variation in population trends and fluctuations, we modelled population trends (μ) and fluctuations (σ2) across all rarity metrics. We conducted one Bayesian linear model per rarity metric per scale (for both global and UK analyses) per population change variable—trends or fluctuations. The response variable was population trend (μ values from state-space models) or population fluctuation (σ2 values from state-space models), and the fixed effects were geographic range (log transformed), mean population size (log transformed) and habitat specificity (number of distinct habitats occupied). The model structures were identical across the different rarity metrics and below we outline the equations for population trends and geographic range:

    $$mu _{i,k,n} = beta _0 + beta _{0,k} + beta _1 ast geographic,range_{i,k,n},$$
    (15)

    $$y_{i,k,n} sim gaussianleft( {mu _{i,k,n},,sigma ^2} right),$$
    (16)

    where geographic rangei,k,n is the logged geographic range of the kth species in the ith time series; β0 and β1 are the global intercept and slope estimate for the geographic range effect (fixed effect), β0j is the species-level departure from β0 (species-level random effect); yi,k,n is the estimate for change in population abundance for the ith population time series from the jth species with the nth geographic range.
    Population trends across species’ IUCN Red List Categories
    To investigate the relationship between-population change and species’ Red List Categories, we modelled population trends (μ) and fluctuations (σ2) as a function of IUCN Red List Categories (categorical variable). We conducted one Bayesian linear model per population change metric per scale (for both global and UK analyses). To test variation in population trends and fluctuations across the types and number of threats to which species are exposed, we conducted a post hoc analysis of trends and fluctuations across threat type (categorical effect) and number of threats that each species is exposed to across its range (in separate models). The model structures were identical to those presented above, except for the fixed effect which was a categorical IUCN Red List Category variable.
    The analytical workflow of our analyses is summarised in conceptual diagrams (Supplementary Figs. 1 and 2) and all code is available on GitHub (https://github.com/gndaskalova/PopChangeRarity, DOI 10.5281/zenodo.3817207).
    Data limitations: taxonomic and geographic gaps
    Our analysis is based on 9286 monitored populations from 2084 species from the largest currently available public database of population time series, the Living Planet Database72. Nevertheless, the data are characterised by both taxonomic and geographic gaps that can influence our findings. For example, there are very few population records from the Amazon and Siberia (Fig. 1b)—two regions currently undergoing rapid environmental changes owing to land-use change and climate change, respectively. In addition, birds represent 63% of all population time series in the Living Planet Database, whilst taxa such as amphibians and sharks where we find declines are included with fewer records (Fig. 2 and Supplementary Fig. 4). On a larger scale, the Living Planet Database under-represents populations outside of Europe and North America and over-represents common and well-studied species62. We found that for the populations and species represented by current monitoring, rarity does not explain variation in population trends, but we note that the relationship between population change and rarity metrics could differ for highly endemic specialist species or species different to the ones included in the Living Planet Database17. As ongoing and future monitoring begins to fill in the taxonomic and geographic gaps in existing datasets, we will be able to reassess and test the generality of the patterns of population change across biomes, taxa, phylogenies, species traits and threats.
    Data limitations: monitoring extent and survey techniques
    The Living Planet Database combines population time series where survey methods were consistent within time series but varied among time series. Thus, among populations, abundance was measured using different units and over varying spatial extents. There are no estimates of error around the raw population abundance values available and detection probability likely varies among species. Thus, it is challenging to make informed decisions about baseline uncertainty in abundance estimates without prior information. We used state-space models to estimate trends and fluctuations to account for these limitations as this modelling framework is particularly appropriate for analyses of data collected using disparate methods41,81,82. Another approach to partially account for observer error that has been applied to the analysis of population trends is the use of occupancy models36. Because the precise coordinates of the polygons where the individual populations were monitored are not available, we were not able to test for the potential confounding effect of monitoring extent, but our sensitivity analysis indicated that survey units do not explain variation in the detected trends (Supplementary Fig. 12).
    Data limitations: temporal gaps
    The population time series we studied cover the period between 1970 and 2014, with both duration of monitoring and the frequency of surveys varying across time series. We omitted populations that had less than five time points of monitoring data, as previous studies of similar population time series data have found that shorter time series are less likely to capture directional trends in abundance63. In a separate analysis, we found significant lags in population change following disturbances (forest loss) and that population monitoring often begins decades to centuries after peak forest loss has occurred at a given site45. The findings of this related study suggest that the temporal span of the population monitoring does not always capture the period of intense environmental change and lags suggest that there might be abundance changes that have not yet manifested themselves. Thus, the detected trends and the baseline across which trends are compared might be influenced by when monitoring takes place and at what temporal frequency. Challenges of analysing time series data are present across not just the Living Planet Database that we analysed, but more broadly across population data in general, including invertebrate datasets65. Nevertheless, the Living Planet Database represents the most comprehensive compilation of vertebrate temporal population records to date, allowing for analyses of the patterns of vertebrate trends and fluctuations around the world.
    Data limitations: time series with low variation
    Eighty populations ( More

  • in

    Larval pesticide exposure impacts monarch butterfly performance

    1.
    Potts, S. G. et al. Global pollinator declines: Trends, impacts and drivers. Trends Ecol. Evol. 25, 345–353 (2010).
    PubMed  Google Scholar 
    2.
    Goulson, D., Nicholls, E., Botías, C. & Rotheray, E. L. Bee declines driven by combined stress from parasites, pesticides, and lack of flowers. Science 347, 1255957 (2015).
    PubMed  Google Scholar 

    3.
    Ollerton, J. Pollinator diversity: Distribution, ecological function, and conservation. Ann. Rev. Ecol. Evol. Syst. 48, 353–376 (2017).
    Google Scholar 

    4.
    Botías, C. et al. Neonicotinoid residues in wildflowers, a potential route of chronic exposure for bees. Environ. Sci. Technol. 49, 12731–12740 (2015).
    ADS  PubMed  Google Scholar 

    5.
    Botías, C., David, A., Hill, E. M. & Goulson, D. Contamination of wild plants near neonicotinoid seed-treated crops, and implications for non-target insects. Sci. Tot. Environ. 566, 269–278 (2016).
    Google Scholar 

    6.
    Long, E. Y. & Krupke, C. H. Non-cultivated plants present a season-long route of pesticide exposure for honey bees. Nat. Commun. 7, 11629 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    7.
    Mogren, C. L. & Lundgren, J. G. Neonicotinoid-contaminated pollinator strips adjacent to cropland reduce honey bee nutritional status. Sci. Rep. 6, 29608 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    8.
    Tsvetkov, N. et al. Chronic exposure to neonicotinoids reduces honey-bee health near corn crops. Science 356, 1395–1397 (2017).
    ADS  CAS  PubMed  Google Scholar 

    9.
    Wood, T. J., Kaplan, I., Zhang, Y. & Szendrei, Z. Honeybee dietary neonicotinoid exposure is associated with pollen collection from agricultural weeds. Proc. R. Soc. B 286, 20190989 (2019).
    CAS  PubMed  Google Scholar 

    10.
    Douglas, M. R. & Tooker, J. F. Large-scale deployment of seed treatments has driven rapid increase in use of neonicotinoid insecticides and preemptive pest management in US field crops. Environ. Sci. Technol. 49, 5088–5097 (2015).
    ADS  CAS  PubMed  Google Scholar 

    11.
    DiBartolomeis, M., Kegley, S., Mineau, P., Radford, R. & Klein, K. An assessment of acute insecticide toxicity loading (AITL) of chemical pesticides used on agricultural land in the United States. PLoS One 14, e0220029 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    12.
    Douglas, M. R., Sponsler, D. B., Lonsdorf, E. V. & Grozinger, C. M. County-level analysis reveals a rapidly shifting landscape of insecticide hazard to honey bees (Apis mellifera) on US farmland. Sci. Rep. 10, 797 (2020).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    13.
    Gilburn, A. S. et al. Are neonicotinoid insecticides driving declines of widespread butterflies?. PeerJ 3, e1402 (2015).
    PubMed  PubMed Central  Google Scholar 

    14.
    Forister, M. L. et al. Increasing neonicotinoid use and the declining butterfly fauna of lowland California. Biol. Lett. 12, 20160475 (2016).
    PubMed  PubMed Central  Google Scholar 

    15.
    Wepprich, T., Adrion, J. R., Ries, L., Wiedmann, J. & Haddad, N. M. Butterfly abundance declines over 20 years of systematic monitoring in Ohio, USA. PLoS One 14, e0216270 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    16.
    Braak, N., Neve, R., Jones, A. K., Gibbs, M. & Breuker, C. J. The effects of insecticides on butterflies: A review. Environ. Pollut. 242, 507–518 (2018).
    CAS  PubMed  Google Scholar 

    17.
    Mulé, R., Sabella, G., Robba, L. & Manachini, B. Systematic review of the effects of chemical insecticides on four common butterfly families. Front. Environ. Sci. 5, 32 (2017).
    Google Scholar 

    18.
    Russell, C. & Schultz, C. B. Effects of grass-specific herbicides on butterflies: An experimental investigation to advance conservation efforts. J. Insect Conserv. 14, 53–63 (2009).
    Google Scholar 

    19.
    Bohnenblust, E., Egan, J. F., Mortensen, D. & Tooker, J. Direct and indirect effects of the synthetic-auxin herbicide dicamba on two lepidopteran species. Environ. Entomol. 42, 586–594 (2013).
    CAS  PubMed  Google Scholar 

    20.
    Hahn, M., Geisthardt, M. & Bruhl, C. A. Effects of herbicide-treated host plants on the development of Mamestra brassicae L. caterpillars. Environ. Toxicol. Chem. 33, 2633–2638 (2014).
    CAS  PubMed  Google Scholar 

    21.
    Stark, J. D., Chen, X. D. & Johnson, C. Effects of herbicides on Behr’s Metalmark butterfly, a surrogate species for the endangered butterfly, Lange’s Metalmark. Environ. Pollut. 164, 24–27 (2012).
    CAS  PubMed  Google Scholar 

    22.
    Eliyahu, D., Applebaum, S. W. & Rafaeli, A. Moth sex-pheromone biosynthesis is inhibited by the herbicide diclofop. Pestic. Biochem. Physiol. 77, 75–81 (2003).
    CAS  Google Scholar 

    23.
    Srivastava, K., Sharma, S., Sharma, D. & Kumar, R. Effect of fungicides on growth and development of Spodoptera litura. Int. J. Life Sci. Sci. Res. 3, 905–908 (2017).
    Google Scholar 

    24.
    Nicodemo, D. et al. Pyraclostrobin impairs energetic mitochondrial metabolism and productive performance of silkworm (Lepidoptera: Bombycidae) caterpillars. J. Econ. Entomol. 111, 1369 (2018).
    PubMed  Google Scholar 

    25.
    Pilling, E. D. & Jepson, P. C. Synergism between EBI fungicides and a pyrethroid insecticide in the honeybee (Apis mellifera). Pestic. Sci. 39, 293–297 (1993).
    CAS  Google Scholar 

    26.
    Pilling, E. D., Bromley-Challenor, K. A. C., Walker, C. H. & Jepson, P. C. Mechanism of synergism between the pyrethroid insecticide λ-cyhalothrin and the imidazole fungicide prochloraz, in the honeybee (Apis mellifera L). Pestic. Biochem. Physiol. 51, 1–11 (1995).
    CAS  Google Scholar 

    27.
    Johnson, R. M., Ellis, M. D., Mullin, C. A. & Frazier, M. Pesticides and honey bee toxicity—USA. Apidologie 41, 312–331 (2010).
    CAS  Google Scholar 

    28.
    Wade, A., Lin, C. H., Kurkul, C., Regan, E. & Johnson, R. M. Combined toxicity of insecticides and fungicides applied to California almond orchards to honey bee larvae and adults. Insects 10, 20 (2019).
    PubMed Central  Google Scholar 

    29.
    Lichtenstein, E. P., Liang, T. T. & Anderegg, B. N. Synergism of insecticides by herbicides. Science 181, 847–849 (1973).
    ADS  CAS  PubMed  Google Scholar 

    30.
    James, D. G. A neonicotinoid insecticide at a rate found in nectar reduces longevity but not oogenesis in monarch butterflies, Danaus plexippus (L.). (Lepidoptera: Nymphalidae). Insects 10, 276 (2019).
    PubMed Central  Google Scholar 

    31.
    Sinha, S. N., Lakhani, K. H. & Davis, B. N. K. Studies on the toxicity of insecticidal drift to the first instar larvae of the large white butterfly Pieris brassicae (Lepidoptera: Pieridae). Ann. Appl. Biol. 116, 27–41 (1990).
    CAS  Google Scholar 

    32.
    Davis, B. N. K., Lakhani, K. H. & Yates, T. J. The hazards of insecticides to butterflies of field margins. Agric. Ecosyst. Environ. 36, 151–161 (1991).
    CAS  Google Scholar 

    33.
    Çilgi, T. & Jepson, P. C. The risks posed by deltamethrin drift to hedgerow butterflies. Environ. Pollut. 87, 1–9 (1995).
    PubMed  Google Scholar 

    34.
    Whitehorn, P. R., Norville, G., Gilburn, A. & Goulson, D. Larval exposure to the neonicotinoid imidacloprid impacts adult size in the farmland butterfly Pieris brassicae. PeerJ 6, e4772 (2018).
    PubMed  PubMed Central  Google Scholar 

    35.
    Agrawal, A. A. & Inamine, H. Mechanisms behind the monarch’s decline. Science 360, 1294–1296 (2018).
    ADS  CAS  PubMed  Google Scholar 

    36.
    Belsky, J. & Joshi, N. K. Assessing role of major drivers in recent decline of monarch butterfly population in North America. Front. Environ. Sci. 6, 86 (2018).
    Google Scholar 

    37.
    Malcolm, S. B. Anthropogenic impacts on mortality and population viability of the monarch butterfly. Ann. Rev. Entomol. 63, 277–302 (2018).
    CAS  Google Scholar 

    38.
    Hartzler, R. G. Reduction in common milkweed (Asclepias syriaca) occurrence in Iowa cropland from 1999 to 2009. Crop Prot. 29, 1542–1544 (2010).
    Google Scholar 

    39.
    Pleasants, J. M. & Oberhauser, K. S. Milkweed loss in agricultural fields because of herbicide use: Effect on the monarch butterfly population. Insect Conserv. Div. 6, 135–144 (2013).
    Google Scholar 

    40.
    Thogmartin, W. E. et al. Monarch butterfly population decline in North America: Identifying the threatening processes. R. Soc. Open. Sci. 4, 170760 (2017).
    ADS  PubMed  PubMed Central  Google Scholar 

    41.
    Thogmartin, W. E. et al. Restoring monarch butterfly habitat in the Midwestern US: ‘all hands on deck’. Environ. Res. Lett. 12, 074005 (2017).
    ADS  Google Scholar 

    42.
    Saunders, S. P., Ries, L., Oberhauser, K. S., Thogmartin, W. E. & Zipkin, E. F. Local and cross-seasonal associations of climate and land use with abundance of monarch butterflies Danaus plexippus. Ecography 41, 278–290 (2018).
    Google Scholar 

    43.
    Stenoien, C. et al. Monarchs in decline: A collateral landscape-level effect of modern agriculture. Insect Sci. 25, 528–541 (2018).
    PubMed  Google Scholar 

    44.
    Krischik, V., Rogers, M., Gupta, G. & Varshney, A. Soil-applied imidacloprid translocates to ornamental flowers and reduces survival of adult Coleomegilla maculata, Harmonia axyridis, and Hippodamia convergens lady beetles, and larval Danaus plexippus and Vanessa cardui butterflies. PLoS One 10, e0119133 (2015).
    PubMed  PubMed Central  Google Scholar 

    45.
    Krishnan, N. et al. Assessing field-scale risks of foliar insecticide applications to monarch butterfly (Danaus plexippus) larvae. Environ. Toxicol. Chem. 39, 923–941 (2020).
    CAS  PubMed  Google Scholar 

    46.
    Pecenka, J. R. & Lundgren, J. G. Non-target effects of clothianidin on monarch butterflies. Sci. Nat. 102, 19 (2015).
    Google Scholar 

    47.
    Bargar, T. A., Hladik, M. L. & Daniels, J. C. Uptake and toxicity of clothianidin to monarch butterflies from milkweed consumption. PeerJ 8, e8669 (2020).
    PubMed  PubMed Central  Google Scholar 

    48.
    Hartzler, R. G. & Buhler, D. D. Occurrence of common milkweed (Asclepias syriaca) in cropland and adjacent areas. Crop Prot. 19, 363–366 (2000).
    Google Scholar 

    49.
    Zaya, D. N., Pearse, I. S. & Spyreas, G. Long-term trends in midwestern milkweed abundances and their relevance to monarch butterfly declines. Bioscience 67, 343–356 (2017).
    Google Scholar 

    50.
    Agrawal, A. A. & Fishbein, M. Plant defense syndromes. Ecology 87, S132–S149 (2006).
    PubMed  Google Scholar 

    51.
    Lefevre, T., Oliver, L., Hunter, M. D. & de Roode, J. C. Evidence for trans-generational medication in nature. Ecol. Lett. 13, 1485–1493 (2010).
    PubMed  Google Scholar 

    52.
    Olaya-Arenas, P. & Kaplan, I. Quantifying pesticide exposure risk for monarch caterpillars on milkweeds bordering agricultural land. Front. Ecol. Evol. 7, 223 (2019).
    Google Scholar 

    53.
    Olaya-Arenas, P., Scharf, M. E. & Kaplan, I. Do pollinators prefer pesticide-free plants? An experimental test with monarchs and milkweeds. J. Appl. Ecol. 20, 20 (2020).
    Google Scholar 

    54.
    Bauerfeind, S. S., Fischer, K., Hartstein, S., Janowitz, S. & Martin-Creuzburg, D. Effects of adult nutrition on female reproduction in a fruit-feeding butterfly: The role of fruit decay and dietary lipids. J. Insect Physiol. 53, 964–973 (2017).
    Google Scholar 

    55.
    Geister, T. L., Lorenz, M. W., Hoffmann, K. H. & Fischer, K. Adult nutrition and butterfly fitness: Effects of diet quality on reproductive output, egg composition, and egg hatching success. Front. Zool. 5, 10 (2008).
    PubMed  PubMed Central  Google Scholar 

    56.
    Van Hook, T., Williams, E. H., Brower, L. P., Borkin, S. & Hein, J. A standardized protocol for ruler-based measurement of wing length in monarch butterflies, Danaus plexippus L. (Nymphalidae, Danainae). Trop. Lep. Res. 22, 42–52 (2012).
    Google Scholar 

    57.
    Davis, A. K. & Holden, M. T. Measuring intraspecific variation in flight-related morphology of monarch butterflies (Danaus plexippus): Which sex has the best flying gear?. J. Insects 20, 591705 (2015).
    Google Scholar 

    58.
    García-Barros, E. Multivariate indices as estimates of dry body weight for comparative study of body size in Lepidoptera. Nota Lepi. 38, 59–74 (2015).
    Google Scholar 

    59.
    Wiesweg, M. Survival Analysis: High-Level Interface for Survival Analysis and Associated Plots. R Package Version 0.1.1. (2019). https://CRAN.R-project.org/package=survivalAnalysis.

    60.
    Wickham, et al. Welcome to the tidyverse. J. Open Source Softw. 4, 1686 (2019).
    ADS  Google Scholar 

    61.
    Kassambara, A. ggpubr: ‘ggplot2’ Based Publication Ready Plots. R Package Version 0.2.5. (2020). https://CRAN.R-project.org/package=ggpubr.

    62.
    Kassambara, A. Rstatix: Pipe-Friendly Framework for Basic Statistical Tests. R package version 0.4.0. (2020). https://CRAN.R-project.org/package=rstatix.

    63.
    Fox, J. & Weisberg, S. An R Companion to Applied Regression 3rd edn. (Sage, Thousand Oaks, 2019).
    Google Scholar 

    64.
    Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer-Verlag, New York, 2016).
    Google Scholar 

    65.
    Kassambara, A. ggcorrplot: Visualization of a Correlation Matrix Using ‘ggplot2’. R Package Version 0.1.3. (2019) https://CRAN.R-project.org/package=ggcorrplot.

    66.
    Hedges, L. V., Gurevitch, J. & Curtis, P. S. The meta-analysis of response ratios in experimental ecology. Ecology 80, 1150–1156 (1999).
    Google Scholar 

    67.
    Schaarschmidt, F. & Gerhard, D. pairwiseCI: Confidence Intervals for Two-Sample Comparisons. R Package Version 0.1-27. (2019) https://CRAN.R-project.org/package=pairwiseCI.

    68.
    Thompson, H. M., Fryday, S. L., Harkin, S. & Milner, S. Potential impacts of synergism in honeybees (Apis mellifera) of exposure to neonicotinoids and sprayed fungicides in crops. Apidologie 45, 545–553 (2014).
    CAS  Google Scholar 

    69.
    Tadei, R. et al. Late effect of larval co-exposure to the insecticide clothianidin and fungicide pyraclostrobin in Africanized Apis mellifera. Sci. Rep. 9, 3277 (2019).
    ADS  PubMed  PubMed Central  Google Scholar 

    70.
    Després, L., David, J. P. & Gallet, C. The evolutionary ecology of insect resistance to plant chemicals. Trends Ecol. Evol. 22, 298–307 (2007).
    PubMed  Google Scholar 

    71.
    Jones, P. L., Petschenka, G., Flacht, L. & Agrawal, A. A. Cardenolide intake, sequestration, and excretion by the monarch butterfly along gradients of plant toxicity and larval ontogeny. J. Chem. Ecol. 45, 264–277 (2019).
    CAS  PubMed  Google Scholar 

    72.
    Karageorgi, M. et al. Genome editing retraces the evolution of toxin resistance in the monarch butterfly. Nature 574, 409–412 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    73.
    Hardy, N. B., Peterson, D. A., Ross, L. & Rosenheim, J. A. Does a plant-eating insect’s diet govern the evolution of insecticide resistance? Comparative tests of the pre-adaptation hypothesis. Evol. Appl. 11, 739–747 (2018).
    PubMed  Google Scholar 

    74.
    Basley, K. & Goulson, D. Effects of field-relevant concentrations of clothianidin on larval development of the butterfly Polyommatus icarus (Lepidoptera, Lycaenidae). Environ. Sci. Technol. 52, 3990–3996 (2018).
    ADS  CAS  PubMed  Google Scholar 

    75.
    Halsch, C. et al. Pesticide contamination of milkweeds across the agricultural, urban, and open spaces of low elevation Northern California. Front. Ecol. Evol. 8, 162 (2020).
    Google Scholar 

    76.
    Altizer, S. & Davis, A. K. Populations of monarch butterflies with different migratory behaviors show divergence in wing morphology. Evolution 64, 1018–1028 (2010).
    PubMed  Google Scholar 

    77.
    Dockx, C. Directional and stabilizing selection on wing size and shape in migrant and resident monarch butterflies, Danaus plexippus (L.). Cuba. Biol. J. Linn. Soc. 92, 605–616 (2007).
    Google Scholar 

    78.
    Satterfield, D. A. & Davis, A. K. Variation in wing characteristics of monarch butterflies during migration: Earlier migrants have redder and more elongated wings. Anim. Migr. 2, 1–7 (2014).
    Google Scholar 

    79.
    Freedman, M. G. & Dingle, H. Wing morphology in migratory North American monarchs: Characterizing sources of variation and understanding changes through time. Anim. Migr. 5, 61–73 (2018).
    Google Scholar 

    80.
    Inamine, H., Ellner, S. P., Springer, J. P. & Agrawal, A. A. Linking the continental migratory cycle of the monarch butterfly to understand its population decline. Oikos 125, 1081–1091 (2016).
    Google Scholar 

    81.
    Nylin, S. & Gotthard, K. Plasticity in life-history traits. Ann. Rev. Entomol. 43, 63–83 (1998).
    CAS  Google Scholar 

    82.
    Awmack, C. S. & Leather, S. R. Host plant quality and fecundity in herbivorous insects. Ann. Rev. Entomol. 47, 817–844 (2002).
    CAS  Google Scholar 

    83.
    Jervis, M. A., Boggs, C. L. & Ferns, P. N. Egg maturation strategy and its associated trade-offs: A synthesis focusing on Lepidoptera. Ecol. Entomol. 30, 359–375 (2005).
    Google Scholar 

    84.
    Oberhauser, K. S. Fecundity, lifespan and egg mass in butterflies: Effects of male-derived nutrients and female size. Funct. Ecol. 11, 166–175 (1997).
    Google Scholar 

    85.
    Main, A. R., Webb, E. B., Goyne, K. W. & Mengel, D. Neonicotinoid insecticides negatively affect performance measures of non-target terrestrial arthropods: A meta-analysis. Ecol. Appl. 28, 1232–1244 (2018).
    PubMed  Google Scholar  More