Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

The spread of resistance to imidacloprid is restricted by thermotolerance in natural populations of Drosophila melanogaster

Abstract

Imidacloprid, the world’s most used insecticide, has caused considerable controversy due to harmful effects on non-pest species and increasing evidence showing that insecticides have become the primary selective force in many insect species. The genetic response to insecticides is heterogeneous across populations and environments, leading to more complex patterns of genetic variation than previously thought. This motivated the investigation of imidacloprid resistance at different temperatures in natural populations of Drosophila melanogaster originating from four climate extremes replicated across two continents. Population and quantitative genomic analysis, supported by functional tests, have revealed a mixed genetic architecture to resistance involving major genes (Paramyosin and Nicotinic-Acetylcholine Receptor Alpha 3) and polygenes with a major trade-off with thermotolerance. Reduced genetic differentiation at resistance-associated loci indicated enhanced gene flow at these loci. Resistance alleles showed stronger evidence of positive selection in temperate populations compared to tropical populations in which chromosomal inversions In(2L)t, In(3R)Mo and In(3R)Payne harbour susceptibility alleles. Polygenic architecture and ecological factors should be considered when developing sustainable management strategies for both pest and beneficial insects.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Sampling locations and effects of imidacloprid resistance and thermotolerance on longevity.
Fig. 2: Genome-wide association for imidacloprid resistance.
Fig. 3: Functional tests for the effect of Prm and nAChRα3.
Fig. 4: Genomic location of candidate genes and genome-wide pattern of diversity and selection.

Similar content being viewed by others

Data availability

Aligned sequence data generated in this study are deposited on NCBI under the BioProject ID PRJNA515537 (https://www.ncbi.nlm.nih.gov/bioproject/515537). All custom codes are deposited on GitHub (https://github.com/aflevel/IMIresist_vs_TEMPtol) and the derived data are deposited on figshare (https://doi.org/10.26188/5b592f305226b).

References

  1. Miles, A. et al. Genetic diversity of the African malaria vector Anopheles gambiae. Nature 552, 96–100 (2017).

    Google Scholar 

  2. Zamoum, T. et al. Does insecticide resistance alone account for the low genetic variability of asexually reproducing populations of the peach-potato aphid Myzus persicae? Heredity 94, 630–639 (2005).

    Article  CAS  Google Scholar 

  3. Woodcock, B. A. et al. Country-specific effects of neonicotinoid pesticides on honey bees and wild bees. Science 356, 1393–1395 (2017).

    Article  CAS  Google Scholar 

  4. Anderson, C. J. et al. Hybridization and gene flow in the mega-pest lineage of moth, Helicoverpa. Proc. Natl Acad. Sci. USA 115, 5034–5039 (2018).

    Article  CAS  Google Scholar 

  5. Daborn, P. J. et al. A single p450 allele associated with insecticide resistance in Drosophila. Science 297, 2253–2256 (2002).

    Article  CAS  Google Scholar 

  6. Pittendrigh, B., Reenan, R., ffrench-Constant, R. H. & Ganetzky, B. Point mutations in the Drosophila sodium channel gene para associated with resistance to DDT and pyrethroid insecticides. Mol. Gen. Genet. 256, 602–610 (1997).

    Article  CAS  Google Scholar 

  7. Fournier, D., Bride, J. M., Hoffmann, F. & Karch, F. Acetylcholinesterase. Two types of modifications confer resistance to insecticide. J. Biol. Chem. 267, 14270–14274 (1992).

    CAS  PubMed  Google Scholar 

  8. Traverso, L. et al. Comparative and functional triatomine genomics reveals reductions and expansions in insecticide resistance-related gene families. PLoS Negl. Trop. Dis. 11, e0005313 (2017).

    Article  Google Scholar 

  9. Faucon, F. et al. Identifying genomic changes associated with insecticide resistance in the dengue mosquito Aedes aegypti by deep targeted sequencing. Genome Res. 25, 1347–1359 (2015).

    Article  CAS  Google Scholar 

  10. Faucon, F. et al. In the hunt for genomic markers of metabolic resistance to pyrethroids in the mosquito Aedes aegypti: an integrated next-generation sequencing approach. PLoS Negl. Trop. Dis. 11, e0005526 (2017).

    Article  Google Scholar 

  11. Ffrench-Constant, R. H. The molecular genetics of insecticide resistance. Genetics 194, 807–815 (2013).

    Article  CAS  Google Scholar 

  12. Messer, P. W. & Petrov, D. A. Population genomics of rapid adaptation by soft selective sweeps. Trends Ecol. Evol. 28, 659–669 (2013).

    Article  Google Scholar 

  13. Ventola, C. L. The antibiotic resistance crisis: part 1: causes and threats. Pharm. Ther. 40, 277–283 (2015).

    Google Scholar 

  14. Tabashnik, B. E. & Carrière, Y. Surge in insect resistance to transgenic crops and prospects for sustainability. Nat. Biotechnol. 35, 926–935 (2017).

    Article  CAS  Google Scholar 

  15. Mallet, J. The evolution of insecticide resistance: have the insects won? Trends Ecol. Evol. 4, 336–340 (1989).

    Article  CAS  Google Scholar 

  16. Munro, A. Economics and biological evolution 1. Environ. Resour. Econ. 9, 429–449 (1997).

    Google Scholar 

  17. Lynd, A. et al. Field, genetic, and modeling approaches show strong positive selection acting upon an insecticide resistance mutation in Anopheles gambiae s.s. Mol. Biol. Evol. 27, 1117–1125 (2010).

    Article  CAS  Google Scholar 

  18. Barnes, K. G. et al. Genomic footprints of selective sweeps from metabolic resistance to pyrethroids in African malaria vectors are driven by scale up of insecticide-based vector control. PLoS Genet. 13, e1006539 (2017).

    Article  Google Scholar 

  19. Jones, C. M. et al. Footprints of positive selection associated with a mutation (N1575Y) in the voltage-gated sodium channel of Anopheles gambiae. Proc. Natl Acad. Sci. USA 109, 6614–6619 (2012).

    Article  CAS  Google Scholar 

  20. Battlay, P., Schmidt, J. M., Fournier-Level, A. & Robin, C. Genomic and transcriptomic associations identify a new insecticide resistance phenotype for the selective sweep at the Cyp6g1 locus of Drosophila melanogaster. Genes Genom. Genet. 6, 2573–2581 (2016).

    CAS  Google Scholar 

  21. Schmidt, J. M. et al. Copy number variation and transposable elements feature in recent, ongoing adaptation at the Cyp6g1 locus. PLoS Genet. 6, e1000998 (2010).

    Article  Google Scholar 

  22. Garud, N. R. et al. Recent selective sweeps in North American Drosophila melanogaster show signatures of soft sweeps. PLoS Genet. 11, e1005004 (2015).

    Article  Google Scholar 

  23. Schmidt, J. M. et al. Insights into DDT resistance from the Drosophila melanogaster genetic reference panel. Genetics 207, 1181–1193 (2017).

    Article  CAS  Google Scholar 

  24. Fournier-Level, A. et al. Behavioural response to combined insecticide and temperature stress in natural populations of Drosophila melanogaster. J. Evol. Biol. 29, 1030–1044 (2016).

    Article  CAS  Google Scholar 

  25. Labbé, P. et al. Forty years of erratic insecticide resistance evolution in the mosquito Culex pipiens. PLoS Genet. 3, e205 (2007).

    Article  Google Scholar 

  26. Andersson, D. I. & Hughes, D. Antibiotic resistance and its cost: is it possible to reverse resistance? Nat. Rev. Microbiol. 8, 260–271 (2010).

    Article  Google Scholar 

  27. Kliot, A. & Ghanim, M. Fitness costs associated with insecticide resistance. Pest Manag. Sci. 68, 1431–1437 (2012).

    Article  CAS  Google Scholar 

  28. Vila-Aiub, M. M., Neve, P. & Powles, S. B. Fitness costs associated with evolved herbicide resistance alleles in plants. New Phytol. 184, 751–767 (2009).

    Article  CAS  Google Scholar 

  29. ffrench-Constant, R. H. & Bass, C. Does resistance really carry a fitness cost? Curr. Opin. Insect Sci. 21, 39–46 (2017).

    Article  Google Scholar 

  30. Wu, C., Davis, A. S. & Tranel, P. J. Limited fitness costs of herbicide-resistance traits in Amaranthus tuberculatus facilitate resistance evolution. Pest Manag. Sci. 74, 293–301 (2018).

    Article  CAS  Google Scholar 

  31. Melnyk, A. H., Wong, A. & Kassen, R. The fitness costs of antibiotic resistance mutations. Evol. Appl. 8, 273–283 (2015).

    Article  Google Scholar 

  32. Bourguet, D., Guillemaud, T., Chevillon, C. & Raymond, M. Fitness costs of insecticide resistance in natural breeding sites of the mosquito Culex pipiens. Evolution 58, 128–135 (2004).

    Article  Google Scholar 

  33. Cousens R. D.. & Fournier-LevelA.. Herbicide resistance costs: what are we actually measuring and why? . Pest Manag. Sci. 74, 1539–1546 (2017).

    Article  Google Scholar 

  34. Nkya, T. E. et al. Insecticide resistance mechanisms associated with different environments in the malaria vector Anopheles gambiae: a case study in Tanzania. Malar. J. 13, 28 (2014).

    Article  Google Scholar 

  35. Sternberg, E. D. & Thomas, M. B. Local adaptation to temperature and the implications for vector-borne diseases. Trends Parasitol. 30, 115–122 (2014).

    Article  Google Scholar 

  36. Lack, J. B., Lange, J. D., Tang, A. D., Corbett-Detig, R. B. & Pool, J. E. A thousand fly genomes: an expanded Drosophila genome nexus. Mol. Biol. Evol. 33, 3308–3313 (2016).

    Article  CAS  Google Scholar 

  37. Hoffmann, A. A., Anderson, A. & Hallas, R. Opposing clines for high and low temperature resistance in Drosophila melanogaster. Ecol. Lett. 5, 614–618 (2002).

    Article  Google Scholar 

  38. Turner, T. L., Levine, M. T., Eckert, M. L. & Begun, D. J. Genomic analysis of adaptive differentiation in Drosophila melanogaster. Genetics 179, 455–473 (2008).

    Article  CAS  Google Scholar 

  39. Kapun, M., Fabian, D. K., Goudet, J. & Flatt, T. Genomic evidence for adaptive inversion clines in Drosophila melanogaster. Mol. Biol. Evol. 33, 1317–1336 (2016).

    Article  CAS  Google Scholar 

  40. Kapun, M., Van Schalkwyk, H., McAllister, B., Flatt, T. & Schlötterer, C. Inference of chromosomal inversion dynamics from Pool-Seq data in natural and laboratory populations of Drosophila melanogaster. Mol. Ecol. 23, 1813–1827 (2014).

    Article  CAS  Google Scholar 

  41. Rane, R. V., Rako, L., Kapun, M., Lee, S. F. & Hoffmann, A. A. Genomic evidence for role of inversion 3RP of Drosophila melanogaster in facilitating climate change adaptation. Mol. Ecol. 24, 2423–2432 (2015).

    Article  Google Scholar 

  42. Matsuda, K. et al. Neonicotinoids: insecticides acting on insect nicotinic acetylcholine receptors. Trends Pharmacol. Sci. 22, 573–580 (2001).

    Article  CAS  Google Scholar 

  43. Venken, K. J. T. et al. MiMIC: a highly versatile transposon insertion resource for engineering Drosophila melanogaster genes. Nat. Methods 8, 737–743 (2011).

    Article  CAS  Google Scholar 

  44. Lavington, E. & Kern, A. D. The effect of common inversion polymorphisms In(2L)t and In(3R)Mo on patterns of transcriptional variation in Drosophila melanogaster. Genes Genom. Genet. 7, 3659–3668 (2017).

    CAS  Google Scholar 

  45. Vörösmarty, C. J. et al. Global threats to human water security and river biodiversity. Nature 467, 555–561 (2010).

    Article  Google Scholar 

  46. Boitard, S., Schlotterer, C., Nolte, V., Pandey, R. V. & Futschik, A. Detecting selective sweeps from pooled next-generation sequencing samples. Mol. Biol. Evol. 29, 2177–2186 (2012).

    Article  CAS  Google Scholar 

  47. Buckingham, S. D., Lapied, B., Le Corronc, H., Grolleau, F. & Sattelle, D. B. Imidacloprid actions on insect neuronal acetylcholine receptors. J. Exp. Biol. 200, 2685–2692 (1997).

    CAS  PubMed  Google Scholar 

  48. Cervera, M., Arredondo, J. J. & Ferreres, R. M. in Nature’s Versatile Engine: Insect Flight Muscle Inside and Out (Springer, New York, 2006); https://doi.org/10.1007/0-387-31213-7_6

  49. Liu, H. et al. Paramyosin phosphorylation site disruption affects indirect flight muscle stiffness and power generation in Drosophila melanogaster. Proc. Natl Acad. Sci. USA 102, 10522–10527 (2005).

    Article  CAS  Google Scholar 

  50. Merzendorfer, H. et al. Genomic and proteomic studies on the effects of the insect growth regulator diflubenzuron in the model beetle species Tribolium castaneum. Insect Biochem. Mol. Biol. 42, 264–276 (2012).

    Article  CAS  Google Scholar 

  51. Zhao, L., Alto, B., Shin, D. & Yu, F. The effect of permethrin resistance on Aedes aegypti transcriptome following ingestion of zika virus infected blood. Viruses 10, 470 (2018).

    Article  Google Scholar 

  52. Gonzalez-Freire, M., de Cabo, R., Studenski, S. A. & Ferrucci, L. The neuromuscular junction: aging at the crossroad between nerves and muscle. Front. Aging Neurosci. 6, 208 (2014).

    Article  Google Scholar 

  53. Takamori, M. Synaptic homeostasis and its immunological disturbance in neuromuscular junction disorders. Int. J. Mol. Sci. 18, E896 (2017).

    Article  Google Scholar 

  54. Joussen, N., Heckel, D. G., Haas, M., Schuphan, I. & Schmidt, B. Metabolism of imidacloprid and DDT by P450 CYP6G1 expressed in cell cultures of Nicotiana tabacum suggests detoxification of these insecticides in Cyp6g1-overexpressing strains of Drosophila melanogaster, leading to resistance. Pest Manag. Sci. 64, 65–73 (2008).

    Article  CAS  Google Scholar 

  55. Fusetto, R., Denecke, S., Perry, T., O’Hair, R. A. J. & Batterham, P. Partitioning the roles of CYP6G1 and gut microbes in the metabolism of the insecticide imidacloprid in Drosophila melanogaster. Sci. Rep. 7, 11339 (2017).

    Article  Google Scholar 

  56. Denecke, S. et al. Multiple P450s and variation in neuronal genes underpins the response to the insecticide imidacloprid in a population of Drosophila melanogaster. Sci. Rep. 7, 11338 (2017).

    Article  Google Scholar 

  57. Broderick, N. A., Raffa, K. F. & Handelsman, J. Midgut bacteria required for Bacillus thuringiensis insecticidal activity. Proc. Natl Acad. Sci. USA 103, 15196–15199 (2006).

    Article  CAS  Google Scholar 

  58. Cheng, D. et al. Gut symbiont enhances insecticide resistance in a significant pest, the oriental fruit fly Bactrocera dorsalis (Hendel). Microbiome 5, 13 (2017).

    Article  Google Scholar 

  59. Dada, N., Sheth, M., Liebman, K., Pinto, J. & Lenhart, A. Whole metagenome sequencing reveals links between mosquito microbiota and insecticide resistance in malaria vectors. Sci. Rep. 8, 2084 (2018).

    Article  Google Scholar 

  60. Kikuchi, Y. et al. Symbiont-mediated insecticide resistance. Proc. Natl Acad. Sci. USA 109, 8618–8622 (2012).

    Article  CAS  Google Scholar 

  61. Hoffmann, A. A. Rapid adaptation of invertebrate pests to climatic stress? Curr. Opin. Insect Sci. 21, 7–13 (2017).

    Article  Google Scholar 

  62. Ayala, D., Ullastres, A. & González, J. Adaptation through chromosomal inversions in Anopheles. Front. Genet. 5, 129 (2014).

    Article  Google Scholar 

  63. Cressey, D. The bitter battle over the world’s most popular insecticides. Nature 551, 156–158 (2017).

    Article  CAS  Google Scholar 

  64. Cernansky, R. Controversial pesticides found in honey samples from six continents. Nature https://doi.org/10.1038/nature.2017.22762 (2017).

  65. Sureda Anfres, M. Controversial insecticides linked to wild bee declines. Nature https://doi.org/10.1038/nature.2016.20446 (2016).

  66. Rundlöf, M. et al. Seed coating with a neonicotinoid insecticide negatively affects wild bees. Nature 521, 77–80 (2015).

    Article  Google Scholar 

  67. Hallmann, C. A., Foppen, R. P. B., van Turnhout, C. A. M., de Kroon, H. & Jongejans, E. Declines in insectivorous birds are associated with high neonicotinoid concentrations. Nature 511, 341–343 (2014).

    Article  CAS  Google Scholar 

  68. Henry, M. et al. A common pesticide decreases foraging success and survival in honey bees. Science 336, 348–350 (2012).

    Article  CAS  Google Scholar 

  69. Hijmans, R. J. & Graham, C. H. The ability of climate envelope models to predict the effect of climate change on species distributions. Glob. Change Biol. 12, 2272–2281 (2006).

    Article  Google Scholar 

  70. Li, H. Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. Preprint at http://arxiv.org/abs/1303.3997 (2013).

  71. Kofler, R., Gómez-Sánchez, D. & Schlötterer, C. PoPoolationTE2: comparative population genomics of transposable elements using pool-seq. Mol. Biol. Evol. 33, 2759–2764 (2016).

    Article  CAS  Google Scholar 

  72. Talevich, E., Shain, A. H., Botton, T. & Bastian, B. C. CNVkit: genome-wide copy number detection and visualization from targeted DNA sequencing. PLoS Comput. Biol. 12, e1004873 (2016).

    Article  Google Scholar 

  73. Fournier-Level, A., Robin, C. & Balding, D. J. GWAlpha: genome-wide estimation of additive effects (alpha) based on trait quantile distribution from pool-sequencing experiments. Bioinformatics 33, 1246–1247 (2016).

    Google Scholar 

  74. Antonov, A. V., Schmidt, E. E., Dietmann, S., Krestyaninova, M. & Hermjakob, H. R spider: a network-based analysis of gene lists by combining signaling and metabolic pathways from Reactome and KEGG databases. Nucleic Acids Res. 38, W78–W83 (2010).

    Article  CAS  Google Scholar 

  75. Ferretti, L., Ramos-Onsins, S. E. & Pérez-Enciso, M. Population genomics from pool sequencing. Mol. Ecol. 22, 5561–5576 (2013).

    Article  Google Scholar 

  76. Kofler, R., Pandey, R. V. & Schlötterer, C. PoPoolation2: identifying differentiation between populations using sequencing of pooled DNA samples (Pool-Seq). Bioinformatics 27, 3435–3436 (2011).

    Article  CAS  Google Scholar 

  77. Boitard, S. et al. Pool-hmm: a Python program for estimating the allele frequency spectrum and detecting selective sweeps from next generation sequencing of pooled samples. Mol. Ecol. Resour. 13, 337–340 (2013).

    Article  Google Scholar 

Download references

Acknowledgements

We thank D. Begun, S. Myles and C. Hart for sharing their laboratory facilities, P. Griffin for fly collection, and K. Charles at Foursights Wines, D. Herbert at Herbert Vineyard, P. Dixon at the Henty Estate, J. Leahy at Becker’s Vineyard, J. Seago at Ponchartrain Vineyard and Landry Vineyards for providing access to their estates. This work was supported by the Human Frontier in Sciences Long-Term fellowship no. LT000907/2012-L awarded to A.F.L.

Author information

Authors and Affiliations

Authors

Contributions

A.F.-L. and C.R. designed the study and wrote the manuscript with contributions from T.P. and A.A.H. A.F.-L., R.T.G. and S.W. performed the experiments. A.F.-L. analysed the data with contributions from A.A.H., R.V.R. and P. Battlay. A.A.H., P. Battlay and T.P. revised the manuscript. A.A.H., M.S., P. Batterham, T.P. and W.C. provided new genetic material.

Corresponding authors

Correspondence to Alexandre Fournier-Level or Charles Robin.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Supplementary Information

Supplementary Figures 1–5, Supplementary Tables 1–5 and 7–9

Reporting Summary

Supplementary Table 6

List of candidate genes for imidacloprid resistance. Candidate genes were identified on the basis of the excess of significantly associated polymorphism within 2.5 kbp of a gene coding sequence in at least two GWAS or present in 300-bp segments with significant copy-number association.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fournier-Level, A., Good, R.T., Wilcox, S.A. et al. The spread of resistance to imidacloprid is restricted by thermotolerance in natural populations of Drosophila melanogaster. Nat Ecol Evol 3, 647–656 (2019). https://doi.org/10.1038/s41559-019-0837-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41559-019-0837-y

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing