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:

Natural selection for imprecise vertical transmission in host–microbiota systems

Subjects

A Publisher Correction to this article was published on 11 March 2022

This article has been updated

Abstract

How and when the microbiome modulates host adaptation remains an evolutionary puzzle, despite evidence that the extended genetic repertoire of the microbiome can shape host phenotypes and fitness. One complicating factor is that the microbiome is often transmitted imperfectly across host generations, leading to questions about the degree to which the microbiome contributes to host adaptation. Here, using an evolutionary model, we demonstrate that decreasing vertical transmission fidelity can increase microbiome variation, and thus phenotypic variation, across hosts. When the most beneficial microbial genotypes change unpredictably from one generation to the next (for example, in variable environments), hosts can maximize fitness by increasing the microbiome variation among offspring, as this improves the chance of there being an offspring with the right microbial combination for the next generation. Imperfect vertical transmission can therefore be adaptive in varying environments. We characterize how selection on vertical transmission is shaped by environmental conditions, microbiome changes during host development and the contribution of other factors to trait variation. We illustrate how environmentally dependent microbial effects can favour intermediate transmission and set our results in the context of examples from natural systems. We also suggest research avenues to empirically test our predictions. Our model provides a basis to understand the evolutionary pathways that potentially led to the wide diversity of microbe transmission patterns found in nature.

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: Natural variation in microbiome transmission fidelity across the eukaryotic kingdom, resulting in variation in microbiome composition across hosts in a population.
Fig. 2: Interactions among transmission fidelity, host phenotypes and fitness, and the environment.
Fig. 3: Conceptual overview of how microbiome transmission could shape host phenotypic variance and when phenotypic variation among genetically similar individuals might be beneficial.
Fig. 4: Environmental predictability and variance together shape the optimal transmission fidelity.
Fig. 5: Including other sources of phenotypic variation among hosts changes selection on transmission fidelity.
Fig. 6: Environmental colonization and within-host proliferation alter host phenotypic variance.

Similar content being viewed by others

Data availability

This study uses computer-generated datasets, which can be created using available R code.

Code availability

An interactive tool to run the model can be found at http://marjoleinbruijning.shinyapps.io/simulhostmicrobiome, and example R code is available on Github: http://github.com/marjoleinbruijning/microbiomeTransmission (https://doi.org/10.5281/zenodo.5534317)84.

Change history

References

  1. Bercik, P. et al. The intestinal microbiota affect central levels of brain-derived neurotropic factor and behavior in mice. Gastroenterology 141, 599–609 (2011).

    Article  CAS  PubMed  Google Scholar 

  2. Johnson, K. V.-A. & Foster, K. R. Why does the microbiome affect behaviour? Nat. Rev. Microbiol. 16, 647–655 (2018).

    Article  CAS  PubMed  Google Scholar 

  3. Sherwin, E., Bordenstein, S. R., Quinn, J. L., Dinan, T. G. & Cryan, J. F. Microbiota and the social brain. Science 366, eaar2016 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Charbonneau, M. R. et al. A microbial perspective of human developmental biology. Nature 535, 48–55 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Blanton, L. V. et al. Gut bacteria that prevent growth impairments transmitted by microbiota from malnourished children. Science 351, aad3311 (2016).

    Article  PubMed  Google Scholar 

  6. Matsuoka, K. & Kanai, T. The gut microbiota and inflammatory bowel disease. Semin. Immunopathol. 37, 47–55 (2015).

    Article  CAS  PubMed  Google Scholar 

  7. Niu, B., Paulson, J. N., Zheng, X. & Kolter, R. Simplified and representative bacterial community of maize roots. Proc. Natl Acad. Sci. USA 114, E2450–E2459 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Berg, M. & Koskella, B. Nutrient- and dose-dependent microbiome-mediated protection against a plant pathogen. Curr. Biol. 28, 2487–2492 (2018).

    Article  CAS  PubMed  Google Scholar 

  9. Wei, Z. et al. Trophic network architecture of root-associated bacterial communities determines pathogen invasion and plant health. Nat. Commun. 6, 8413 (2015).

  10. Keebaugh, E. S., Yamada, R., Obadia, B., Ludington, W. B. & William, W. J. Microbial quantity impacts Drosophila nutrition, development, and lifespan. iScience 4, 247–259 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Camarinha-Silva, A. et al. Host genome influence on gut microbial composition and microbial prediction of complex traits in pigs. Genetics 206, 1637–1644 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Difford, G. F. et al. Host genetics and the rumen microbiome jointly associate with methane emissions in dairy cows. PLoS Genet. 14, e1007580 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Moran, N. A. & Sloan, D. B. The hologenome concept: helpful or hollow? PLoS Biol. 13, e1002311 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Henry, L. P., Bruijning, M., Forsberg, S. K. G. & Ayroles, J. F. The microbiome extends host evolutionary potential. Nat. Commun. 12, 5141 (2021).

  15. Foster, K. R., Schluter, J., Coyte, K. Z. & Rakoff-Nahoum, S. The evolution of the host microbiome as an ecosystem on a leash. Nature 548, 43–51 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Baumann, P. Biology of bacteriocyte-associated endosymbionts of plant sap-sucking insects. Annu. Rev. Microbiol. 59, 155–189 (2005).

    Article  CAS  PubMed  Google Scholar 

  17. Douglas, A. E. Nutritional interactions in insect–microbial symbioses: aphids and their symbiotic bacteria Buchnera. Annu. Rev. Entomol. 43, 17–37 (1998).

    Article  CAS  PubMed  Google Scholar 

  18. Roughgarden, J., Gilbert, S. F., Rosenberg, E., Zilber-Rosenberg, I. & Lloyd, E. A. Holobionts as units of selection and a model of their population dynamics and evolution. Biol. Theory 13, 44–65 (2018).

    Article  Google Scholar 

  19. Fukatsu, T. & Hosokawa, T. Capsule-transmitted gut symbiotic bacterium of the Japanese common plataspid stinkbug, Megacopta punctatissima. Appl. Environ. Microbiol. 68, 389–396 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Kaiwa, N. et al. Symbiont-supplemented maternal investment underpinning host’s ecological adaptation. Curr. Biol. 24, 2465–2470 (2014).

    Article  CAS  PubMed  Google Scholar 

  21. Jahnes, B. C., Herrmann, M. & Sabree, Z. L. Conspecific coprophagy stimulates normal development in a germ-free model invertebrate. PeerJ 7, e6914 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Estes, A. M. et al. Brood ball-mediated transmission of microbiome members in the dung beetle, Onthophagus taurus (Coleoptera: Scarabaeidae). PLoS ONE 8, e79061 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. van Vliet, S. & Doebeli, M. The role of multilevel selection in host microbiome evolution. Proc. Natl Acad. Sci. USA 116, 20591–20597 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Zeng, Q., Wu, S., Sukumaran, J. & Rodrigo, A. Models of microbiome evolution incorporating host and microbial selection. Microbiome 5, 127 (2017).

  25. Björk, J. R., Diez-Vives, C., Astudillo-Garcia, C., Archie, E. A. & Montoya, J. M. Vertical transmission of sponge microbiota is inconsistent and unfaithful. Nat. Ecol. Evol. 3, 1172–1183 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Douglas, A. E. & Werren, J. H. Holes in the hologenome: why host–microbe symbioses are not holobionts. mBio 7, e02099-15 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Hammer, T. J. & Moran, N. A. Links between metamorphosis and symbiosis in holometabolous insects. Phil. Trans. R. Soc. B 374, 20190068 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Metcalf, C. J. E., Henry, L. P., Rebolleda-Gomez, M. & Koskella, B. Why evolve reliance on the microbiome for timing of ontogeny?. mBio 10, e01496-19 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Bruijning, M., Metcalf, C. J. E., Jongejans, E. & Ayroles, J. F. The evolution of variance control. Trends Ecol. Evol. 35, 22–23 (2020).

    Article  PubMed  Google Scholar 

  30. Bull, J. J. Evolution of phenotypic variance. Evolution 41, 303–315 (1987).

    Article  CAS  PubMed  Google Scholar 

  31. Philippi, T. & Seger, J. Hedging one’s evolutionary bets, revisited. Trends Ecol. Evol. 4, 41–44 (1989).

    Article  CAS  PubMed  Google Scholar 

  32. Vasseur, D. A. & Yodzis, P. The color of environmental noise. Ecology 85, 1146–1152 (2004).

    Article  Google Scholar 

  33. Halley, J. M. Ecology, evolution and 1f-noise. Trends Ecol. Evol. 11, 33–37 (1996).

    Article  CAS  PubMed  Google Scholar 

  34. Botero, C. A., Weissing, F. J., Wright, J. & Rubenstein, D. R. Evolutionary tipping points in the capacity to adapt to environmental change. Proc. Natl Acad. Sci. USA 112, 184–189 (2015).

    Article  CAS  PubMed  Google Scholar 

  35. Burns, A. R. et al. Contribution of neutral processes to the assembly of gut microbial communities in the zebrafish over host development. ISME J. 10, 655–664 (2016).

    Article  CAS  PubMed  Google Scholar 

  36. Kolodny, O. et al. Coordinated change at the colony level in fruit bat fur microbiomes through time. Nat. Ecol. Evol. 3, 116–124 (2019).

    Article  PubMed  Google Scholar 

  37. Sieber, M. et al. Neutrality in the metaorganism. PLoS Biol. 17, e3000298 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Burns, A. R. et al. Interhost dispersal alters microbiome assembly and can overwhelm host innate immunity in an experimental zebrafish model. Proc. Natl Acad. Sci. USA 114, 11181–11186 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Moeller, A. H., Suzuki, T. A., Phifer-Rixey, M. & Nachman, M. W. Transmission modes of the mammalian gut microbiota. Science 362, 453–457 (2018).

    Article  CAS  PubMed  Google Scholar 

  40. Zapién-Campos, R., Sieber, M. & Traulsen, A. Stochastic colonization of hosts with a finite lifespan can drive individual host microbes out of equilibrium. PLoS Comput. Biol. 16, e1008392 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  41. De Vries, E. J., Jacobs, G., Sabelis, M. W., Menken, S. B. J. & Breeuwer, J. A. J. Diet-dependent effects of gut bacteria on their insect host: the symbiosis of Erwinia sp. and western flower thrips. Proc. R. Soc. Lond. B 271, 2171–2178 (2004).

    Article  Google Scholar 

  42. Johnson, N. C., Graham, J. H. & Smith, F. A. Functioning of mycorrhizal associations along the mutualism–parasitism continuum. N. Phytol. 135, 575–585 (1997).

    Article  Google Scholar 

  43. Cheney, K. L. & Côté, I. M. Mutualism or parasitism? The variable outcome of cleaning symbioses. Biol. Lett. 1, 162–165 (2005).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Russell, J. A. & Moran, N. A. Costs and benefits of symbiont infection in aphids: variation among symbionts and across temperatures. Proc. R. Soc. B 273, 603–610 (2006).

    Article  PubMed  Google Scholar 

  45. Oliver, K. M., Degnan, P. H., Burke, G. R. & Moran, N. A. Facultative symbionts in aphids and the horizontal transfer of ecologically important traits. Annu. Rev. Entomol. 55, 247–266 (2010).

    Article  CAS  PubMed  Google Scholar 

  46. Oliver, K. M., Russell, J. A., Moran, N. A. & Hunter, M. S. Facultative bacterial symbionts in aphids confer resistance to parasitic wasps. Proc. Natl Acad. Sci. USA 100, 1803–1807 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Oliver, K. M., Campos, J., Moran, N. A. & Hunter, M. S. Population dynamics of defensive symbionts in aphids. Proc. R. Soc. B 275, 293–299 (2008).

    Article  PubMed  Google Scholar 

  48. Ives, A. R. et al. Self-perpetuating ecological–evolutionary dynamics in an agricultural host–parasite system. Nat. Ecol. Evol. 4, 702–711 (2020).

    Article  PubMed  Google Scholar 

  49. Chen, D.-Q., Montllor, C. B. & Purcell, A. H. Fitness effects of two facultative endosymbiotic bacteria on the pea aphid, Acyrthosiphon pisum, and the blue alfalfa aphid, A. kondoi. Entomol. Exp. Appl. 95, 315–323 (2000).

    Article  Google Scholar 

  50. Montllor, C. B., Maxmen, A. & Purcell, A. H. Facultative bacterial endosymbionts benefit pea aphids Acyrthosiphon pisum under heat stress. Ecol. Entomol. 27, 189–195 (2002).

    Article  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Kikuchi, Y. & Yumoto, I. Efficient colonization of the bean bug Riptortus pedestris by an environmentally transmitted Burkholderia symbiont. Appl. Environ. Microbiol. 79, 2088–2091 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Gould, A. L. et al. Microbiome interactions shape host fitness. Proc. Natl Acad. Sci. USA 115, E11951–E11960 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Ellner, S. P. & Rees, M. Integral projection models for species with complex demography. Am. Nat. 167, 410–428 (2006).

    Article  PubMed  Google Scholar 

  55. Caswell, H. Matrix Population Models: Construction, Analysis and Interpretation (Sinauer Associates, 2001).

  56. Asnicar, F. et al. Studying vertical microbiome transmission from mothers to infants by strain-level metagenomic profiling. mSystems 2, e00164-16 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Yassour, M. et al. Strain-level analysis of mother-to-child bacterial transmission during the first few months of life. Cell Host Microbe 24, 146–154 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Ferretti, P. et al. Mother-to-infant microbial transmission from different body sites shapes the developing infant gut microbiome. Cell Host Microbe 24, 133–145 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Nyholm, S. V. & McFall-Ngai, M. The winnowing: establishing the squid-Vibrio symbiosis. Nat. Rev. Microbiol. 2, 632–642 (2004).

    Article  CAS  PubMed  Google Scholar 

  60. Kikuchi, Y., Hosokawa, T. & Fukatsu, T. Insect–microbe mutualism without vertical transmission: a stinkbug acquires a beneficial gut symbiont from the environment every generation. Appl. Environ. Microbiol. 73, 4308–4316 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Ibáñez, F., Tonelli, M. L., Muñoz, V., Figueredo, M. S. & Fabra, A. in Endophytes: Biology and Biotechnology (ed. Maheshwari, D.) 25–40 (Springer, 2017).

  62. Werren, J. H., Baldo, L. & Clark, M. E. Wolbachia: master manipulators of invertebrate biology. Nat. Rev. Microbiol. 6, 741–751 (2008).

    Article  CAS  PubMed  Google Scholar 

  63. Teixeira, L., Ferreira, Á. & Ashburner, M. The bacterial symbiont Wolbachia induces resistance to RNA viral infections in Drosophila melanogaster. PLoS Biol. 6, 2753–2763 (2008).

    Article  CAS  Google Scholar 

  64. Chrostek, E. et al. Wolbachia variants induce differential protection to viruses in Drosophila melanogaster: a phenotypic and phylogenomic analysis. PLoS Genet. 9, e1003896 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Chrostek, E. & Teixeira, L. Mutualism breakdown by amplification of Wolbachia genes. PLoS Biol. 13, e1002065 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Ravel, C., Michalakis, Y. & Charmet, G. The effect of imperfect transmission on the frequency of mutualistic seed-borne endophytes in natural populations of grasses. Oikos 80, 18–24 (1997).

    Article  Google Scholar 

  67. Buskirk, S. W., Rokes, A. B. & Lang, G. I. Adaptive evolution of nontransitive fitness in yeast. eLife 9, e62238 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Clune, J. et al. Natural selection fails to optimize mutation rates for long-term adaptation on rugged fitness landscapes. PLoS Comput. Biol. 4, e1000187 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  69. King, O. D. & Masel, J. The evolution of bet-hedging adaptations to rare scenarios. Theor. Popul. Biol. 72, 560–575 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Liu, X.-D., Lei, H.-X. & Chen, F.-F. Infection pattern and negative effects of a facultative endosymbiont on its insect host are environment-dependent. Sci. Rep. 9, 4013 (2019).

  71. Oyserman, B. O. et al. Extracting the GEMs: genotype, environment, and microbiome interactions shaping host phenotypes. Front. Microbiol. 11, 3444 (2021).

    Article  Google Scholar 

  72. Rock, D. I. et al. Context-dependent vertical transmission shapes strong endosymbiont community structure in the pea aphid, Acyrthosiphon pisum. Mol. Ecol. 27, 2039–2056 (2018).

    Article  PubMed  Google Scholar 

  73. Osaka, R., Nomura, M., Watada, M. & Kageyama, D. Negative effects of low temperatures on the vertical transmission and infection density of a Spiroplasma endosymbiont in Drosophila hydei. Curr. Microbiol. 57, 335–339 (2008).

    Article  CAS  PubMed  Google Scholar 

  74. Gundel, P. E. et al. Imperfect vertical transmission of the endophyte Neotyphodium in exotic grasses in grasslands of the Flooding Pampa. Microb. Ecol. 57, 740 (2009).

    Article  PubMed  Google Scholar 

  75. Li, L. & Ma, Z. S. Testing the neutral theory of biodiversity with human microbiome datasets. Sci. Rep. 6, 31448 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Foster, K. R. & Bell, T. Competition, not cooperation, dominates interactions among culturable microbial species. Curr. Biol. 22, 1845–1850 (2012).

    Article  CAS  PubMed  Google Scholar 

  77. Sprockett, D., Fukami, T. & Relman, D. A. Role of priority effects in the early-life assembly of the gut microbiota. Nat. Rev. Gastroenterol. Hepatol. 15, 197–205 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  78. Stein, R. R. et al. Ecological modeling from time-series inference: insight into dynamics and stability of intestinal microbiota. PLoS Comput. Biol. 9, e1003388 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  79. Scheuring, I. & Yu, D. W. How to assemble a beneficial microbiome in three easy steps. Ecol. Lett. 15, 1300–1307 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Roughgarden, J. Holobiont evolution: Mathematical model with vertical vs. horizontal microbiome transmission. Phil. Theory Pract. Biol. 12, 002 (2020).

  81. Theis, K. R. et al. Getting the hologenome concept right: an eco-evolutionary framework for hosts and their microbiomes. mSystems 1, e00028-16 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Sloan, W. T. et al. Quantifying the roles of immigration and chance in shaping prokaryote community structure. Environ. Microbiol. 8, 732–740 (2006).

    Article  PubMed  Google Scholar 

  83. Gillespie, J. Polymorphism in random environments. Theor. Popul. Biol. 4, 193–195 (1973).

    Article  Google Scholar 

  84. Bruijning, M. Code for: Natural selection for imprecise vertical transmission in host-microbiota systems. Zenodo https://doi.org/10.5281/zenodo.5534317 (2021).

  85. Sauer, C., Dudaczek, D., Hölldobler, B. & Gross, R. Tissue localization of the endosymbiotic bacterium “Candidatus Blochmannia floridanus” in adults and larvae of the carpenter ant Camponotus floridanus. Appl. Environ. Microbiol. 68, 4187–4193 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Koga, R., Meng, X.-Y., Tsuchida, T. & Fukatsu, T. Cellular mechanism for selective vertical transmission of an obligate insect symbiont at the bacteriocyte–embryo interface. Proc. Natl Acad. Sci. USA 109, E1230–E1237 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Brentassi, M. E. et al. Bacteriomes of the corn leafhopper, Dalbulus maidis (DeLong & Wolcott, 1923) (Insecta, Hemiptera, Cicadellidae: Deltocephalinae) harbor Sulcia symbiont: molecular characterization, ultrastructure, and transovarial transmission. Protoplasma 254, 1421–1429 (2017).

    Article  CAS  PubMed  Google Scholar 

  88. Picazo, D. R. et al. Horizontally transmitted symbiont populations in deep-sea mussels are genetically isolated. ISME J. 13, 2954–2968 (2019).

    Article  Google Scholar 

  89. Gilbert, J. A. et al. Current understanding of the human microbiome. Nat. Med. 24, 392–400 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Korpela, K. et al. Selective maternal seeding and environment shape the human gut microbiome. Genome Res. 28, 561–568 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Walters, W. A. et al. Large-scale replicated field study of maize rhizosphere identifies heritable microbes. Proc. Natl Acad. Sci. USA 115, 7368–7373 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  92. Douglas, A. E. Simple animal models for microbiome research. Nat. Rev. Microbiol. 17, 764–775 (2019).

    Article  CAS  PubMed  Google Scholar 

  93. Sommer, F. et al. The gut microbiota modulates energy metabolism in the hibernating brown bear Ursus arctos. Cell Rep. 14, 1655–1661 (2016).

    Article  CAS  PubMed  Google Scholar 

  94. David, L. A. et al. Diet rapidly and reproducibly alters the human gut microbiome. Nature 505, 559–563 (2014).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

M.B. is supported by NWO Rubicon grant no. 019.192EN.017, C.J.E.M. by National Science Foundation grant no. 1753993 and J.F.A. by National Institutes of Health grant no. GM124881.

Author information

Authors and Affiliations

Authors

Contributions

M.B. and L.P.H. conceived the ideas. M.B. developed the modelling framework with strong input from L.P.H., S.K.G.F., C.J.E.M. and J.F.A. M.B. wrote the first draft, and all authors contributed substantially to revisions.

Corresponding author

Correspondence to Marjolein Bruijning.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Ecology and Evolution thanks Kevin Foster and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Increased stochasticity in small host populations.

In Fig. 2, we show how the microbiome transmission fidelity shapes host phenotype distributions. In doing so, we simulated large populations, consisting of 500 individuals, in order to obtain robust results. This results in a limited role of stochasticity, explaining the relatively low variation across replicated simulations (see shaded regions in Fig. 2a,b). In smaller populations, however, populations are, unsurprisingly, more sensitive to stochastic processes. Here, we set transmission fidelity τ at 1, implying strict vertical transmission, and assessed the average deviation from P = 0 for varying population sizes. In small populations, there is an increase in the number of maladapted populations (that is a larger deviation from the optimal phenotype). Grey dots indicate individual simulations (30 per population size), red lines indicate median values for each population size. τ = 1; ω2 = 1; \(\sigma _\varphi ^2\)=2; Vα=0.01.

Extended Data Fig. 2 Results when the average phenotype is kept constant.

Plots show relationship between transmission fidelity and phenotypic variance (upper row), deviation from the long-term optimal mean (middle row) and long-term fitness (bottom row) when selection can shift the mean phenotype (left; corresponds to Fig. 2) or when keeping the mean phenotype fixed at 0 (right). This was done by mean-centering phenotypes in each time step, by subtracting each phenotype by the average time-specific phenotype. When we keep the mean host phenotype at 0, we can use Bull’s modeling framework30 to calculate long-term fitness (crosses in bottom right panel), based on the relation between transmission fidelity and phenotypic variance (Appendix S2).

Extended Data Fig. 3 The difference between heritability and inheritance.

Heritability (averaged across 5 replicated simulations) is a function of the transmission fidelity, colonization from the environment, and the number of microbial generations within a host generation. Heritability is measured as the slope of a regression between parent and offspring phenotypes upon the moment of reproduction, averaged across time steps. If there is only one microbial generation within a host generation and/or without colonization, the heritability equals the transmission fidelity. However, when one or both increase, heritability decreases, illustrating the difference between inheritance and heritability.

Extended Data Fig. 4 Microbial variation among hosts within a host generation.

Within one host generation, increased colonization from the microbial source pool decreases microbial variation among hosts, as predicted from metacommunity theory. Variation among hosts is calculated as the average microbial diversity within each host, divided by the total microbial diversity across all hosts.

Extended Data Fig. 5 Empirical approaches to testing how microbiome transmission can affect host fitness.

a) Soil transfers in plant microbiomes can be used to enforce strict environmental acquisition or vertical transmission. By either successively inoculating plant generations with their initial starting microbial community (upper row in panel A-i), or passaging the microbial community from the previous to the next generation (bottom row in panel A-i), transmission of microbes can be controlled. Each host generation, artificial selection can be used to select plants based on their phenotype (for example plant size, illustrated here), whereby selection regimes vary (imposing either constant or fluctuating selection). Based on our results, we expect that under constant selection, strict vertical transmission increases fitness compared to strict horizontal transmission, as it allows phenotype distributions to respond to selection. In contrast, under sufficiently large fluctuating selection, vertical transmission reduces phenotypic variation, decreasing long-term fitness. b) A single microbe with a clear effect on host performance can also be used to study selection on transmission fidelity. As discussed in the manuscript, aphid fitness effects of several vertically transmitted symbionts, as well as their environmental-dependence, are quite well understood45. This makes aphids arguably a suitable system to study selection on vertical transmission fidelity. To do so, one could vary the symbiont frequency in different aphid populations (panel B-i). Populations can be followed through time, while keeping symbiont frequencies constant. Based on our results, we expect that under constant selection, a symbiont frequency of 100% (or 0%) optimizes population growth (panel B-ii), which can be realized by perfect vertical transmission. Under fluctuating selection, some intermediate symbiont frequency might be favored (panel B-ii), which can be achieved by noisy vertical transmission.

Extended Data Fig. 6 Schematic overview of our model.

Upper panel shows the different steps of our simulations, and the parameters that we vary. Bottom panel shows the output that we obtain from each simulation.

Extended Data Fig. 7 The evolution of transmission fidelity in mixed populations, under different environmental conditions.

For the analyses presented in our main text, we assessed long-term fitness of each strategy (transmission fidelity) separately, and take the strategy with the highest long-term fitness (calculated as the geometric mean) for what would evolve in a mixed population. Here, we simulated dynamics of mixed populations (consisting of 5000 individuals), and show that this yield the same outcomes. We assigned to each host a random transmission fidelity at the start of a simulation run, and performed 5 replicated runs for each environmental condition. Three bottom plots show how the composition of transmission fidelities in a host population changes over time in specific simulation runs (letters A-C corresponding to scenarios depicted in upper graph).

Supplementary information

Supplementary Information

Supplementary Information 1 and 2.

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bruijning, M., Henry, L.P., Forsberg, S.K.G. et al. Natural selection for imprecise vertical transmission in host–microbiota systems. Nat Ecol Evol 6, 77–87 (2022). https://doi.org/10.1038/s41559-021-01593-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41559-021-01593-y

This article is cited by

Search

Quick links

Nature Briefing Anthropocene

Sign up for the Nature Briefing: Anthropocene newsletter — what matters in anthropocene research, free to your inbox weekly.

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