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.

  • Letter
  • Published:

Transmission of mitochondrial mutations and action of purifying selection in Drosophila melanogaster

Subjects

Abstract

It is not known how selection affects mutations in the multiple copies of the mitochondrial genome1,2,3,4,5,6,7,8,9,10,11. We transferred cytoplasm between D. melanogaster embryos carrying mitochondrial mutations to create heteroplasmic lines transmitting two mitochondrial genotypes. Increased temperature imposed selection against a temperature-sensitive mutation affecting cytochrome oxidase, driving decreases in the abundance of the mutant genome over successive generations. Selection did not influence the health or fertility of the flies but acted during midoogenesis to influence competition between the genomes. Mitochondria might incur an advantage through selective localization, survival or proliferation, yet timing and insensitivity to park mutation suggest that preferential proliferation underlies selection. Selection drove complete replacement of the temperature-sensitive mitochondrial genome by a wild-type genome but also stabilized the multigenerational transmission of two genomes carrying complementing detrimental mutations. While they are so balanced, these stably transmitted mutations have no detrimental phenotype, but their segregation could contribute to disease phenotypes and somatic aging.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Stable transmission of genetically marked mitochondrial genomes in the background of mt:ND2del1.
Figure 2: Segregation of mitochondrial genotypes in progeny.
Figure 3: Purifying selection against the temperature-sensitive genome.
Figure 4: Temperature shift of mothers followed by analysis of mitochondrial genome abundance in eggs shows the timing of selection in the germ line.

Similar content being viewed by others

References

  1. Chinnery, P.F. et al. The inheritance of mitochondrial DNA heteroplasmy: random drift, selection or both? Trends Genet. 16, 500–505 (2000).

    Article  CAS  Google Scholar 

  2. Fan, W. et al. A mouse model of mitochondrial disease reveals germline selection against severe mtDNA mutations. Science 319, 958–962 (2008).

    Article  CAS  Google Scholar 

  3. Stewart, J.B. et al. Strong purifying selection in transmission of mammalian mitochondrial DNA. PLoS Biol. 6, e10–e71 (2008).

    Article  Google Scholar 

  4. Freyer, C. et al. Variation in germline mtDNA heteroplasmy is determined prenatally but modified during subsequent transmission. Nat. Genet. 44, 1282–1285 (2012).

    Article  CAS  Google Scholar 

  5. Volz-Lingenhöhl, A., Solignac, M. & Sperlich, D. Stable heteroplasmy for a large-scale deletion in the coding region of Drosophila subobscura mitochondrial DNA. Proc. Natl. Acad. Sci. USA 89, 11528–11532 (1992).

    Article  Google Scholar 

  6. Hauswirth, W.W. & Laipis, P.J. Mitochondrial DNA polymorphism in a maternal lineage of Holstein cows. Proc. Natl. Acad. Sci. USA 79, 4686–4690 (1982).

    Article  CAS  Google Scholar 

  7. Jenuth, J.P., Peterson, A.C., Fu, K. & Shoubridge, E.A. Random genetic drift in the female germline explains the rapid segregation of mammalian mitochondrial DNA. Nat. Genet. 14, 146–151 (1996).

    Article  CAS  Google Scholar 

  8. Cree, L.M. et al. A reduction of mitochondrial DNA molecules during embryogenesis explains the rapid segregation of genotypes. Nat. Genet. 40, 249–254 (2008).

    Article  CAS  Google Scholar 

  9. Cao, L. et al. New evidence confirms that the mitochondrial bottleneck is generated without reduction of mitochondrial DNA content in early primordial germ cells of mice. PLoS Genet. 5, e1000756 (2009).

    Article  Google Scholar 

  10. Cao, L. et al. The mitochondrial bottleneck occurs without reduction of mtDNA content in female mouse germ cells. Nat. Genet. 39, 386–390 (2007).

    Article  CAS  Google Scholar 

  11. Wai, T., Teoli, D. & Shoubridge, E.A. The mitochondrial DNA genetic bottleneck results from replication of a subpopulation of genomes. Nat. Genet. 40, 1484–1488 (2008).

    Article  CAS  Google Scholar 

  12. Matsuura, E.T., Chigusa, S.I. & Niki, Y. Induction of mitochondrial DNA heteroplasmy by intra- and interspecific transplantation of germ plasm in Drosophila. Genetics 122, 663–667 (1989).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Xu, H., DeLuca, S.Z. & O'Farrell, P.H. Manipulating the metazoan mitochondrial genome with targeted restriction enzymes. Science 321, 575–577 (2008).

    Article  CAS  Google Scholar 

  14. DeLuca, S.Z. & O'Farrell, P.H. Barriers to male transmission of mitochondrial DNA in sperm development. Dev. Cell 22, 660–668 (2012).

    Article  CAS  Google Scholar 

  15. Solignac, M., Génermont, J., Monnerot, M. & Mounolou, J.C. Drosophila mitochondrial genetics: evolution of heteroplasmy through germ line cell divisions. Genetics 117, 687–696 (1987).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Solignac, M.G., Nermont, J., Monnerot, M. & Mounolou, J.C. Genetics of mitochondria in Drosophila: mtDNA inheritance in heteroplasmic strains of D. mauritiana. Mol. Gen. Genet. 197, 183–188 (1984).

    Article  CAS  Google Scholar 

  17. King, R.C. Ovarian Development in Drosophila melanogaster (Academic Press, New York, 1970).

  18. Hill, J.H., Chen, Z. & Xu, H. Selective propagation of functional mitochondrial DNA during oogenesis restricts the transmission of a deleterious mitochondrial variant. Nat. Genet. 10.1038/ng.2920 (9 March 2014).

  19. Cox, R.T. & Spradling, A.C. A Balbiani body and the fusome mediate mitochondrial inheritance during Drosophila oogenesis. Development 130, 1579–1590 (2003).

    Article  CAS  Google Scholar 

  20. Wallace, D.C. Why do we still have a maternally inherited mitochondrial DNA? Insights from evolutionary medicine. Annu. Rev. Biochem. 76, 781–821 (2007).

    Article  CAS  Google Scholar 

  21. Ogienko, A.A., Fedorova, S.A. & Baricheva, E.M. Basic aspects of ovarian development in Drosophila melanogaster. Genetika 43, 1341–1357 (2007).

    CAS  PubMed  Google Scholar 

  22. Robinson, D.N., Cant, K. & Cooley, L. Morphogenesis of Drosophila ovarian ring canals. Development 120, 2015–2025 (1994).

    CAS  PubMed  Google Scholar 

  23. Horne-Badovinac, S. & Bilder, D. Mass transit: epithelial morphogenesis in the Drosophila egg chamber. Dev. Dyn. 232, 559–574 (2005).

    Article  CAS  Google Scholar 

  24. Wright, S. Evolution and the Genetics of Populations. Volume II: Theory of Gene Frequencies (University of Chicago Press, Chicago, 1984).

Download references

Acknowledgements

We thank D. Martinez for helping perform the park RNAi experiment and G. Mardon (Baylor College of Medicine) for kindly providing the dpkΔ21 strain. This research was supported by the US National Institutes of Health (GM086854 and ES020725) and by United Mitochondrial Disease Foundation funding to P.H.O. H.M. was supported by the Human Frontiers Science Program (LT000138/2010-l).

Author information

Authors and Affiliations

Authors

Contributions

H.M., H.X. and P.H.O. designed the research, H.M. performed the research, H.X. contributed reagents, and H.M. and P.H.O. wrote the manuscript.

Corresponding author

Correspondence to Patrick H O'Farrell.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Abundance distribution of donor mtDNA in the progeny (n = 316) of a single mother (mother 1 in Table 1) changed over successive days.

Transmission at 22 °C from a mother carrying 1.29% of her mtDNA as mt:CoIT300I in the background of mt:ND2del1. (A) Histograms showing the number of progeny within different abundance intervals for the mt:CoIT300I genome, where each histogram shows the progeny arising from the eggs collected over the indicated time period. The black bar indicates the number of progeny lacking PCR-detectable mt:CoIT300I genomes. (B) Means, percentages of homoplasmic individuals (P0), variances and estimated effective inherited units (N) for serial cohorts of progeny. See Supplementary Table 3 for predictions of the number of segregating units using different formulas.

Supplementary Figure 2 Elimination of the temperature-sensitive genome after multiple generations of selection at 29 °C.

The abundance of the temperature-sensitive mutant was measured by PCR amplifying a mtDNA region (mt1579–2799) using mtDNA from 30 adults as template followed by restriction digestion using XhoI in 4 heteroplasmic lines. (A) The upper panel shows that the mt:ND2del1 + mt:CoIT300I genome, when coexisting with wild type, declined to a few percent after ten generations of selection. The lower panel is Southern blot analysis of the digested PCR products, which showed disappearance of the temperature-sensitive allele in the tested heteroplasmic lines at generation 18. (B) mt:ND2del1 + mt:CoIT300I double-mutant genome was eliminated when coexisting with mt:ND2del1 in 2 of the 4 lines after 18 generations of selection.

Supplementary Figure 3 Biased transmission mt:ND2del1 when coresident with the wild-type or mt:CoIR301Q genome at 25 °C.

(A) The abundance of the wild-type genome was followed in five heteroplasmic lines for six generations. (B) The abundance of mt:CoIR301Q was followed in four heteroplasmic lines for ten generations (note that the amount of mt:CoIR301Q was not measured at generation 6– 9).

Supplementary Figure 4 Probing the level of selection.

(A) Abundance of the temperature-sensitive genome affected neither mtDNA copy number nor fecundity. Individual mothers were born and raised at 29 °C. (B) Selection does not increase as the number of germline stem cell divisions increases in the mother. The abundance of mt:CoIT300I is shown for three heteroplasmic mothers (mt:CoIT300I/mt:ND2del1 females raised at 29 °C) and the eggs she laid over successive days. A reduction in abundance is seen between the mother and eggs, and there is no significant change in the degree of this reduction over time.

Supplementary Figure 5 Temperature-dependent changes in the proportions of the mt:CoIT300I genome over generations in additional lines.

(A) Eight mt:CoIT300I/mt:ND2del1 sublines at 29 °C (restrictive) and (B) ten mt:CoIT300I/mt:ND2del1 sublines at 22 °C (permissive). All 18 sublines had more than 40% defective genome to start with.

Supplementary Figure 6 Schematic of the establishment of heteroplasmic lines.

Original heteroplasmic mothers were generated through pole plasm transplantation, and their individual F1 female progeny (known as G0 or founding mothers) were crossed to mt:ND2del1 males to establish isofemale lines. Sometimes, single females from further generations were used to establish sublines.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6 and Supplementary Tables 1–5 (PDF 1394 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ma, H., Xu, H. & O'Farrell, P. Transmission of mitochondrial mutations and action of purifying selection in Drosophila melanogaster. Nat Genet 46, 393–397 (2014). https://doi.org/10.1038/ng.2919

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.2919

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