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:

A yeast living ancestor reveals the origin of genomic introgressions

Abstract

Genome introgressions drive evolution across the animal1, plant2 and fungal3 kingdoms. Introgressions initiate from archaic admixtures followed by repeated backcrossing to one parental species. However, how introgressions arise in reproductively isolated species, such as yeast4, has remained unclear. Here we identify a clonal descendant of the ancestral yeast hybrid that founded the extant Saccharomyces cerevisiae Alpechin lineage5, which carries abundant Saccharomyces paradoxus introgressions. We show that this clonal descendant, hereafter defined as a ‘living ancestor’, retained the ancestral genome structure of the first-generation hybrid with contiguous S. cerevisiae and S. paradoxus subgenomes. The ancestral first-generation hybrid underwent catastrophic genomic instability through more than a hundred mitotic recombination events, mainly manifesting as homozygous genome blocks generated by loss of heterozygosity. These homozygous sequence blocks rescue hybrid fertility by restoring meiotic recombination and are the direct origins of the introgressions present in the Alpechin lineage. We suggest a plausible route for introgression evolution through the reconstruction of extinct stages and propose that genome instability allows hybrids to overcome reproductive isolation and enables introgressions to emerge.

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: The ancestor of the Alpechin lineage.
Fig. 2: LOH blocks guide recombination of the subgenomes of the hybrid.
Fig. 3: Reconstructing the origin of introgressions.
Fig. 4: The asexual fitness landscape.

Similar content being viewed by others

Data availability

The genome sequences generated in this study are available at Sequence Read Archive (SRA), NCBI under the accession codes BioProject ID PRJNA594913, Biosample ID SAMN13540515–SAMN13540586. Re-analysed sequencing data are available as part of the project ERP014555 in the Sequence Read Archive. The phenotype data are available within the Supplementary Information files. All the strains generated in this work are freely available upon request. Source data are provided with this paper.

Code availability

Custom Bash and Perl scripts used in this study are available at: https://github.com/mdangiolo89/A-yeast-living-ancestor-reveals-the-origin-of-genomic-introgressions

References

  1. Edelman, N. B. et al. Genomic architecture and introgression shape a butterfly radiation. Science 366, 594–599 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  2. Arnold, B. J. et al. Borrowed alleles and convergence in serpentine adaptation. Proc. Natl Acad. Sci. USA 113, 8320–8325 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Sun, Y. et al. Large-scale introgression shapes the evolution of the mating-type chromosomes of the filamentous ascomycete Neurospora tetrasperma. PLoS Genet. 8, e1002820 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Greig, D. Reproductive isolation in Saccharomyces. Heredity 102, 39–44 (2009).

    Article  CAS  PubMed  Google Scholar 

  5. Peter, J. et al. Genome evolution across 1,011 Saccharomyces cerevisiae isolates. Nature 556, 339–344 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  6. Wolf, A. B. & Akey, J. M. Outstanding questions in the study of archaic hominin admixture. PLoS Genet. 14, e1007349–e14 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Harrison, R. G. & Larson, E. L. Hybridization, introgression, and the nature of species boundaries. J. Hered. 105 (Suppl. 1), 795–809 (2014).

    Article  PubMed  Google Scholar 

  8. Mallet, J. Hybridization as an invasion of the genome. Trends Ecol. Evol. 20, 229–237 (2005).

    Article  PubMed  Google Scholar 

  9. Martin, S. H. & Jiggins, C. D. Interpreting the genomic landscape of introgression. Curr. Opin. Genet. Dev. 47, 69–74 (2017).

    Article  CAS  PubMed  Google Scholar 

  10. Morales, L. & Dujon, B. Evolutionary role of interspecies hybridization and genetic exchanges in yeasts. Microbiol. Mol. Biol. Rev. 76, 721–739 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Marsit, S. et al. Evolutionary biology through the lens of budding yeast comparative genomics. Nat. Rev. Genet. 18, 581–598 (2017).

    Article  CAS  PubMed  Google Scholar 

  12. Hunter, N., Chambers, S. R., Louis, E. J. & Borts, R. H. The mismatch repair system contributes to meiotic sterility in an interspecific yeast hybrid. EMBO J. 15, 1726–1733 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Kao, K. C., Schwartz, K. & Sherlock, G. A genome-wide analysis reveals no nuclear dobzhansky-muller pairs of determinants of speciation between S. cerevisiae and S. paradoxus, but suggests more complex incompatibilities. PLoS Genet. 6, e1001038 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Greig, D., Louis, E. J., Borts, R. H. & Travisano, M. Hybrid speciation in experimental populations of yeast. Science 298, 1773–1775 (2002).

    Article  ADS  CAS  PubMed  Google Scholar 

  15. Liti, G., Barton, D. B. H. & Louis, E. J. Sequence diversity, reproductive isolation and species concepts in Saccharomyces. Genetics 174, 839–850 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Liti, G. et al. Population genomics of domestic and wild yeasts. Nature 458, 337–341 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  17. Almeida, P. et al. A Gondwanan imprint on global diversity and domestication of wine and cider yeast Saccharomyces uvarum. Nat. Commun. 5, 4044 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  18. Barbosa, R. et al. Evidence of natural hybridization in Brazilian wild lineages of Saccharomyces cerevisiae. Genome Biol. Evol. 8, 317–329 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Gallone, B. et al. Interspecific hybridization facilitates niche adaptation in beer yeast. Nat. Ecol. Evol. 3, 1562–1575 (2019).

    Article  PubMed  Google Scholar 

  20. Langdon, Q. K. et al. Fermentation innovation through complex hybridization of wild and domesticated yeasts. Nat. Ecol. Evol. 3, 1576–1586 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Lee, H.-Y. et al. Incompatibility of nuclear and mitochondrial genomes causes hybrid sterility between two yeast species. Cell 135, 1065–1073 (2008).

    Article  CAS  PubMed  Google Scholar 

  22. Pontes, A., Čadež, N., Gonçalves, P. & Sampaio, J. P. A quasi-domesticate relic hybrid population of Saccharomyces cerevisiae × S. paradoxus adapted to olive brine. Front. Genet. 10, 449 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Bergström, A. et al. A high-definition view of functional genetic variation from natural yeast genomes. Mol. Biol. Evol. 31, 872–888 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Yue, J.-X. et al. Contrasting evolutionary genome dynamics between domesticated and wild yeasts. Nat. Genet. 49, 913–924 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Symington, L. S., Rothstein, R. & Lisby, M. Mechanisms and regulation of mitotic recombination in Saccharomyces cerevisiae. Genetics 198, 795–835 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Dutta, A. et al. Genome dynamics of hybrid Saccharomyces cerevisiae during vegetative and meiotic divisions. G3 (Bethesda) 7, 3669–3679 (2017).

    Article  CAS  Google Scholar 

  27. Tattini, L. et al. Accurate tracking of the mutational landscape of diploid hybrid genomes. Mol. Biol. Evol. 36, 2861–2877 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Dunn, B. et al. Recurrent rearrangement during adaptive evolution in an interspecific yeast hybrid suggests a model for rapid introgression. PLoS Genet. 9, e1003366 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Smukowski Heil, C. S. et al. Loss of heterozygosity drives adaptation in hybrid yeast. Mol. Biol. Evol. 34, 1596–1612 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Bozdag, G. O. et al. Engineering recombination between diverged yeast species reveals genetic incompatibilities. Preprint at https://doi.org/10.1101/755165 (2019).

  31. Hou, J., Friedrich, A., de Montigny, J. & Schacherer, J. Chromosomal rearrangements as a major mechanism in the onset of reproductive isolation in Saccharomyces cerevisiae. Curr. Biol. 24, 1153–1159 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Rogers, D. W., McConnell, E., Ono, J. & Greig, D. Spore-autonomous fluorescent protein expression identifies meiotic chromosome mis-segregation as the principal cause of hybrid sterility in yeast. PLoS Biol. 16, e2005066 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Laureau, R. et al. Extensive recombination of a yeast diploid hybrid through meiotic reversion. PLoS Genet. 12, e1005781 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. McClure, A. W., Jacobs, K. C., Zyla, T. R. & Lew, D. J. Mating in wild yeast: delayed interest in sex after spore germination. Mol. Biol. Cell 29, 3119–3127 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Kaniewski, D. et al. Primary domestication and early uses of the emblematic olive tree: palaeobotanical, historical and molecular evidence from the Middle East. Biol. Rev. Camb. Philos. Soc. 87, 885–899 (2012).

    Article  PubMed  Google Scholar 

  36. Greig, D., Borts, R. H., Louis, E. J. & Travisano, M. Epistasis and hybrid sterility in Saccharomyces. Proc. Biol. Sci. 269, 1167–1171 (2002).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Charron, G., Marsit, S., Hénault, M., Martin, H. & Landry, C. R. Spontaneous whole-genome duplication restores fertility in interspecific hybrids. Nat. Commun. 10, 4126–10 (2019).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  38. Tsai, I. J., Bensasson, D., Burt, A. & Koufopanou, V. Population genomics of the wild yeast Saccharomyces paradoxus: quantifying the life cycle. Proc. Natl Acad. Sci. USA 105, 4957–4962 (2008).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  39. Magwene, P. M. et al. Outcrossing, mitotic recombination, and life-history trade-offs shape genome evolution in Saccharomyces cerevisiae. Proc. Natl Acad. Sci. USA 108, 1987–1992 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  40. Huerta-Sánchez, E. et al. Altitude adaptation in Tibetans caused by introgression of Denisovan-like DNA. Nature 512, 194–197 (2014).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  41. Enard, D. & Petrov, D. A. Evidence that RNA viruses drove adaptive introgression between Neanderthals and modern humans. Cell 175, 360–371.e13 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Prüfer, K. et al. The complete genome sequence of a Neanderthal from the Altai Mountains. Nature 505, 43–49 (2014).

    Article  ADS  PubMed  CAS  Google Scholar 

  43. Slon, V. et al. The genome of the offspring of a Neanderthal mother and a Denisovan father. Nature 561, 113–116 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  44. Liti, G., Warringer, J. & Blomberg, A. Isolation and laboratory domestication of natural yeast strains. Cold Spring Harb. Protoc. 2017, pdb.prot089052–6 (2017).

    Article  PubMed  Google Scholar 

  45. Haber, J. E. Mating-type genes and MAT switching in Saccharomyces cerevisiae. Genetics 191, 33–64 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Vázquez-García, I. et al. Clonal heterogeneity influences the fate of new adaptive mutations. Cell Rep. 21, 732–744 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. De Chiara, M. et al. Domestication reprogrammed the budding yeast life cycle. Preprint at https://doi.org/10.1101/2020.02.08.939314 (2020).

  48. Beltran, G., Novo, M., Rozès, N., Mas, A. & Guillamón, J. M. Nitrogen catabolite repression in Saccharomyces cerevisiae during wine fermentations. FEMS Yeast Res. 4, 625–632 (2004).

    Article  CAS  PubMed  Google Scholar 

  49. Zackrisson, M. et al. Scan-o-matic: high-resolution microbial phenomics at a massive scale. G3 (Bethesda) 6, 3003–3014 (2016).

    Article  CAS  Google Scholar 

  50. Barré, B. et al. Intragenic repeat expansion in the cell wall protein gene HPF1 controls yeast chronological aging. Genome Res. 30, 697–710 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Simpson, J. T. et al. ABySS: a parallel assembler for short read sequence data. Genome Res. 19, 1117–1123 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Yue, J.-X. & Liti, G. Long-read sequencing data analysis for yeasts. Nat. Protocols 13, 1213–1231 (2018).

    Article  CAS  PubMed  Google Scholar 

  53. Koren, S. et al. Canu: scalable and accurate long-read assembly via adaptive k-mer weighting and repeat separation. Genome Res. 27, 722–736 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Chin, C.-S. et al. Nonhybrid, finished microbial genome assemblies from long-read SMRT sequencing data. Nat. Methods 10, 563–569 (2013).

    Article  CAS  PubMed  Google Scholar 

  55. Walker, B. J. et al. Pilon: an integrated tool for comprehensive microbial variant detection and genome assembly improvement. PLoS ONE 9, e112963 (2014).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  56. Kurtz, S. et al. Versatile and open software for comparing large genomes. Genome Biol. 5, R12 (2004).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Lechner, M. et al. Proteinortho: detection of (co-)orthologs in large-scale analysis. BMC Bioinformatics 12, 124 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  58. McLaren, W. et al. The Ensembl variant effect predictor. Genome Biol. 17, 122 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  59. Fay, J. C. & Benavides, J. A. Evidence for domesticated and wild populations of Saccharomyces cerevisiae. PLoS Genet. 1, 66–71 (2005).

    Article  CAS  PubMed  Google Scholar 

  60. Rolland, T. & Dujon, B. Yeasty clocks: dating genomic changes in yeasts. C. R. Biol. 334, 620–628 (2011).

    Article  CAS  PubMed  Google Scholar 

  61. Bradley, R. K. et al. Fast statistical alignment. PLOS Comput. Biol. 5, e1000392 (2009).

    Article  MathSciNet  PubMed  PubMed Central  CAS  Google Scholar 

  62. Slater, G. S. C. & Birney, E. Automated generation of heuristics for biological sequence comparison. BMC Bioinformatics 6, 31 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  63. Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  64. Bouckaert, R. et al. BEAST 2.5: An advanced software platform for Bayesian evolutionary analysis. PLOS Comput. Biol. 15, e1006650 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Hasegawa, M., Kishino, H. & Yano, T. Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. J. Mol. Evol. 22, 160–174 (1985).

    Article  CAS  PubMed  ADS  Google Scholar 

  66. Drummond, A. J. & Suchard, M. A. Bayesian random local clocks, or one rate to rule them all. BMC Biol. 8, 114 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Heled, J. & Drummond, A. J. Bayesian inference of species trees from multilocus data. Mol. Biol. Evol. 27, 570–580 (2010).

    Article  CAS  PubMed  Google Scholar 

  68. Shen, X.-X. et al. Tempo and mode of genome evolution in the budding yeast subphylum. Cell 175, 1533–1545.e20 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Pan, J. et al. A hierarchical combination of factors shapes the genome-wide topography of yeast meiotic recombination initiation. Cell 144, 719–731 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Ruderfer, D. M., Pratt, S. C., Seidel, H. S. & Kruglyak, L. Population genomic analysis of outcrossing and recombination in yeast. Nat. Genet. 38, 1077–1081 (2006).

    Article  CAS  PubMed  Google Scholar 

  71. Gerke, J., Lorenz, K. & Cohen, B. Genetic interactions between transcription factors cause natural variation in yeast. Science 323, 498–501 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  72. Roncoroni, M. et al. The yeast IRC7 gene encodes a β-lyase responsible for production of the varietal thiol 4-mercapto-4-methylpentan-2-one in wine. Food Microbiol. 28, 926–935 (2011).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank A. Bergström, G. Fischer, B. Llorente, S. Mozzachiodi and L. Tattini for critical reading of the manuscript. This work was supported by Agence Nationale de la Recherche (ANR-11-LABX-0028-01, ANR-13-BSV6-0006-01, ANR-15-IDEX-01, ANR-16-CE12-0019 and ANR-18-CE12-0004) and the Swedish Research Council (2014-6547, 2014-4605 and 2018-03638). We thank the Institut Curie NGS platform for generating the sequence data. J.S. is a fellow of the University of Strasbourg Institute for Advanced Study and a member of the Institut Universitaire de France.

Author information

Authors and Affiliations

Authors

Contributions

M.D., M.D.C., J.X.Y., J.W. and G.L. designed the experiments; M.D., M.D.C. and J.X.Y. performed the genomic analyses; M.D., A.I. and A.L. performed and analysed sporulation efficiency, spore viability and LOH rates; S.S. and K.P. performed and analysed the growth phenotype data; B.B. performed and analysed the chronological life span; E.G., R.M., J.S., J.W. and G.L. contributed with resources and reagents; J.W. and G.L. conceived and supervised the project; M.D., J.W. and G.L. wrote the paper.

Corresponding author

Correspondence to Gianni Liti.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks the 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 figures and tables

Extended Data Fig. 1 Genome-wide distribution of LOH and introgressions.

a, Genotype maps of the genome of the living ancestor (top) and the 17 Alpechins. The DNA of the living ancestor is coloured to denote S. cerevisiae (blue) and S. paradoxus (red) DNA, respectively. The DNA of the Alpechins is coloured to denote the S. cerevisiae background (blue) and the interspersed introgressions that are homozygous (red) or heterozygous (grey) for S. paradoxus DNA. There is extensive overlap between the positions of S. paradoxus LOH in the living ancestor and introgressions in the Alpechins. b, Magnified view of chromosome II. Colours as in a. White circles indicate centromeres. Genomic coordinates are based on the S. cerevisiae DBVPG6765 genome assembly.

Source data

Extended Data Fig. 2 Alternative evolutionary models of the origins of the living ancestor and Alpechins.

The genome-wide overlap between LOH, introgressions and mutated sites within LOH regions with two copies of S. cerevisiae or S. paradoxus DNA allowed us to test the alternative models 1 (left) and 2 (right). See Supplementary Discussion 1 for details. Top boxes report the number of introgressions and LOH, partitioned into overlapping and non-overlapping events. Bottom boxes report the number of de novo point mutations in LOH blocks of the living ancestor. Left and right boxes show mutations in S. paradoxus and S. cerevisiae LOH blocks, respectively. Numbers around the top chromosome pair report the total number of de novo mutations, partitioned into those occurring on the S. cerevisiae and S. paradoxus haplotype, respectively. Numbers below the bottom chromosome pair report de novo mutations that are shared between the living ancestor and the Alpechins, again partitioned into those occurring on the S. cerevisiae and S. paradoxus haplotype, respectively. Haplotype phasing data are available only for the living ancestor. Patterns of overlap between S. paradoxus LOH/introgressions and shared de novo mutations between the living ancestor’s S. paradoxus LOH and the Alpechin strains are compatible with model 1. By contrast, the sequence of events in model 2 makes it incompatible with the observed data: the emergence of S. cerevisiae LOH and their de novo mutations after the hybridization between an Alpechin and a S. paradoxus implies that none of these de novo mutations should be present in any Alpechin strain. The presence of 20 mutations that are shared between the Alpechins and the living ancestor rules out model 2 in favour of model 1.

Source data

Extended Data Fig. 3 Genome structure of the living ancestor.

Each coloured bar represents a chromosome, with dark tones representing S. paradoxus DNA and light tones representing S. cerevisiae DNA. Dark blocks within S. cerevisiae chromosomes represent S. paradoxus LOH, and light blocks within S. paradoxus chromosomes represent S. cerevisiae LOH. White circles indicate centromeres and numbers (from 1 to 5) annotate the recombination breakpoints of the gross chromosomal rearrangements listed in the top-right box. Subgenome assignment for rearranged chromosomes is based on centromere flanking sequences. Genomic coordinates of S. cerevisiae and S. paradoxus chromosomes are based on the genome assemblies of DBVPG6765 and CBS432 strains, respectively. We cannot confidently say whether intrinsic (genetic) or extrinsic (environmentally induced) genome instability explain the astounding genome instability signature, nor we can say whether it emerged from a single event or accumulated gradually.

Extended Data Fig. 4 LOH size distribution.

a, Histograms of LOH sizes in the living ancestor. Left, the size histogram of S. cerevisiae LOH events (two copies of S. cerevisiae DNA retained and loss of the S. paradoxus DNA copy). Bin width is 4 kb. Right, the size histogram of the S. paradoxus LOHs (two copies of S. paradoxus DNA retained and loss of the S. cerevisiae DNA copy). Dark and light colours respectively indicate terminal and interstitial LOH events. Insets show magnified views of the x axis up to 30 kb. Bin width is 2.5 kb. b, Size histograms of new LOH events in the mutation accumulation lines. Bin width is 15 kb. Insets show magnified views of the x axis up to 30 kb. Bin width is 2 kb. Panel order and colour codes are as in a.

Source data

Extended Data Fig. 5 Genome-wide genotype of the living ancestor gametes.

a, We used the genome sequences of 25 living ancestor haploid gametes to infer their recombination landscape. Red and blue colours represent S. paradoxus and S. cerevisiae DNA, respectively. The top part shows the living ancestor. Grey segments represent heterozygous regions for which both S. paradoxus and S. cerevisiae DNA is present and correspond to duplicated regions derived from segregation of rearranged chromosomes or aneuploidies. Aneuploidies were all gain-of-copies, with chromosome III particularly prone to duplicate (7 out of 25 gametes), consistent with the presence of a single LOH block encompassing the centromere, where crossovers are rare69. Chromosomes with intra-homologue and inter-homologue gross chromosomal rearrangements are indicated by an asterisk and a plus sign, respectively. Genomic coordinates are based on the S. cerevisiae DBVPG6765 genome assembly. b, Percentage of the genome bearing S. cerevisiae or S. paradoxus DNA in the 25 living ancestor gametes. Colours as in a. c, The number of observed S. cerevisiae (blue) and S. paradoxus (red) haplotype blocks in three living ancestor gametes. These blocks derive from the combination of meiotic recombination events and pre-existing LOH. As comparison, we calculated the theoretical number of haplotype blocks expected from a single meiosis, considering only the number of crossovers, in an experimental S. cerevisiaeS. paradoxus F1 hybrid (data from ref. 13, brown bar), and in the living ancestor if LOH were not present (green bar). We estimated the number of expected haplotypes (y) as y = 16(x + 1), in which x represents the number of crossovers per chromosome. d, Size distribution of S. paradoxus haplotype blocks in three living ancestor gametes (top) and of introgression blocks in three Alpechins (bottom).

Source data

Extended Data Fig. 6 Living ancestor clonal evolution.

a, Genome-wide genotype maps of the living ancestor and its diploid clones derived by mutation accumulation line protocol (MAL) through 120 single-cell bottlenecks and return-to-growth clones derived by inducing and then aborting meiosis. Grey represents heterozygous DNA, red represents LOH blocks with two S. paradoxus DNA copies, and blue represents LOH blocks with two S. cerevisiae DNA copies. Green arrows indicate new LOH recombination events, purple arrows indicate chromosome homogenization by loss and re-synthesis events. Genomic coordinates are based on the S. cerevisiae DBVPG6765 genome assembly. b, c, Sporulation efficiency (at day 3) (b) and spore viability (c) of the living ancestor, its mutation accumulation lines (n = 8), its mitotically growing clones (T0, n = 6) and return-to-growth clones (n = 12). Box plots are as in Fig. 2. Right, spore viability of each single living ancestor’s mutation accumulation line. One line (A887R29) has a huge increase in spore viability owing to tetraploidization. The pink dotted line indicates the spore viability of the living ancestor. d, Coverage plots of chromosome XI of a tetraploid (A887R29) and a diploid (A887R30) mutation accumulation lines. The plots show a balanced chromosome number for the A887R30 diploid clone (1 copy of S. cerevisiae and 1 copy of S. paradoxus) and an unbalanced one for the A887R29 tetraploid clone (3 copies of S. cerevisiae and 1 copy of S. paradoxus). Dots represent mean coverage values in 10-kb non-overlapping windows. Lines represent whole-chromosome coverage medians.

Source data

Extended Data Fig. 7 Reconstruction of the hybridization-to-introgression model.

a, Grey-dotted box contains a representation of the classic model for the origin of introgressions. The discovery of the living ancestor rewrote part of this mechanistic process and enabled us to test the downstream stages. For each stage of the model, we indicate the measured median gamete viability (percentage). The initial S. cerevisiaeS. paradoxus F1 hybrid derived from an outbreeding event and this has been shown to be rare38,70, thus one or very few hybridizations initiated the process. Next, the F1 hybrid expanded clonally to a possibly large population size. This mitotic clonal expansion was followed by the genome instability and LOH formation phase. Genome instability of such magnitude might be rare and produce unviable combinations, probably reducing the ancestral hybrid population with LOH to one or few individuals. Next, the hybrid with LOH expanded mitotically, generating a population that kept evolving, for example, by accumulating additional LOH. Part of the population was maintained as asexual descendant, whereas part went through sexual reproduction and produced the Alpechin lineage. It is likely that the sporulation step was at the population level, therefore producing many spores. Next, one or more spores backcrossed to S. cerevisiae, which might occur rarely since it requires outbreeding. We propose that this step bottlenecked the population to few individuals, consistent with our data showing that Alpechin strains share many of the LOH blocks. The living ancestor gametes were more fertile with S. cerevisiae than with S. paradoxus. We also crossed three living ancestor gametes to a haploid derivative of an Alpechin strain (CPG) and found fertility to be higher than in backcrosses with the S. cerevisiae lineage lacking introgressions. This is consistent with the Alpechin lineage being closer to the living ancestor. b, Sporulation efficiency after 3 days in sporulating conditions of hybridization-to-introgression stages. Box plots are as in Fig. 2. n = biologically independent samples. Sc, n = 1; Sp, n = 2; F1 hybrid, n = 2; LA, n = 1; LA gametes, n = 25; Bc Sc, n = 8; Bc Sp, n = 8; Alpechin (Alp) n = 16; LA gametes × Alp, n = 3; Sc × Alp, n = 1; Sp × Alp, n = 1. The S. cerevisiae × Alpechin cross is a slow sporulator, consistent with the presence of inefficient sporulation alleles71. c, Sporulation efficiency (at day 3) of living ancestor gametes grouped by number of aneuploidies. Box plots are as in Fig. 2. 0 aneuploidies (n = 5), 1 aneuploidy (n = 6), 2 aneuploidies (n = 12), 3 aneuploidies (n = 1), 4 aneuploidies (n = 1). Pink dots indicate spores with an additional copy of chromosome III, which remained haploid and thus do not sporulate. d, Spore viability of hybridization-to-introgression stages. Box plots are as in Fig. 2. Sc, n = 1; Sp, n = 2; F1 hybrid, n = 2; LA, n = 1; LA gametes, n = 13; Bc Sc, n = 8; Bc Sp, n = 8; Alp, n = 4; LA gametes × Alp, n = 3; Sc × Alp, n = 1; Sp × Alp, n = 1. e, Percentage of viable spores for haploid-stable living ancestor gametes backcrossed to S. cerevisiae (blue line boxes) or S. paradoxus (red line boxes), grouped by their number of aneuploidies. Box plots are as in Fig. 2. n = biologically independent samples. 0 aneuploidies (n = 6), 1 aneuploidy (n = 6), 2 aneuploidies (n = 2), 3 aneuploidies (n = 2).

Source data

Extended Data Fig. 8 Mapping the slow growth phenotype in the Alpechin AQA strain.

a, Percentage of introgressed DNA in the Alpechin strains ranked for amount of heterozygous introgressions (grey, ordered first in the stack bar). Most introgressed material is in homozygous state (red, ordered second). Three strains, including AQA, have a considerable fraction of heterozygous introgressions. The rest of the genome up to 100% is S. cerevisiae background (blue, ordered third) and the y axis is cut at 10%. b, Tetrad dissection of the Alpechin AQA strain showing a 2:2 colony size segregation pattern (two slow-growing and two normal-growing colonies in each meiosis), consistent with a single heterozygous variant underlying the growth defect. c, Each dot represents the allele frequency (AF) difference of reference and non-reference alleles across the genome between pools of big and small spores. Two regions show fixation of the opposite alleles in the big and small pools and chromosome XII is compatible with the assumption that the deleterious allele associated with the small phenotype has the lowest frequency in the S. cerevisiae population. d, Chromosome XII allele frequency zoom and map of homozygous (red) and heterozygous (grey) introgressions on a S. cerevisiae genome backbone (blue) of the Alpechin strain AQA. The growth defect maps to the SFI1 gene, which is far from the closest introgression, indicating that the slow growth phenotype is not driven by a deleterious effect of heterozygous introgressions masked by the S. cerevisiae allele. The white circle indicates the centromere. Genomic coordinates are based on the S. cerevisiae DBVPG6765 genome assembly.

Source data

Extended Data Fig. 9 The competitive fitness of the living ancestor.

a, Mitotic growth (absolute doubling time, h) of the living ancestor (n = 40) as compared to the closest extant relatives of its S. cerevisiae DBVPG6765 (top left, n = 8) and S. paradoxus N17 (top middle, n = 16) parents and a diploid experimental F1 hybrid between the two (top right, n = 16), the living ancestor’s diploid gametes (bottom left, n = 200) and backcrosses of its haploid-stable gametes to S. cerevisiae (bottom middle, n = 64) and S. paradoxus parents (bottom right, n = 64). The number of observations (n) results from the combination of the strains in each group and the number of replicates performed in the experiment. Significant differences enumerated in the main text are indicated by asterisks or name (two-tailed Welsh test, FDR q = 0.05) and presented in Supplementary Table 7. For the living ancestor gametes, we instead indicate (red text) the rare environments where no significant difference was found. b, The percentage of surviving cells (mean ± s.d.) for each step of the hybridization-to-introgression process during a 25-day chronological lifespan experiment. n = biologically independent samples. S. cerevisiae (n = 1), S. paradoxus (n = 2), experimental first-generation (F1) hybrid (n = 2), living ancestor (n = 1), living ancestor gametes (n = 25), backcross to S. cerevisiae (n = 8), backcross to S. paradoxus (n = 8), Alpechin (n = 16). c, Gating strategy used for the estimation of the percentage of surviving cells shown in b.

Source data

Extended Data Fig. 10 The adaptive value of introgressions.

a, Heat map of the relative doubling time (mean of n = 40 replicates for the living ancestor and n = 8 replicates for all the other strains) of the living ancestor, its modern Alpechin descendants and the closest living relatives of its S. cerevisiae and S. paradoxus parents. Strains were grouped (left) on the basis of doubling time similarity using hierarchical clustering (Pearson’s r, complete linkage) and branch colours highlight four major clusters. Environments in which the living ancestor grows significantly (two-tailed Welsh test, FDR q = 0.05) faster (green) or slower (red) than the Alpechins are coloured. b, Difference in relative doubling time between Alpechins (n = 128) and S. cerevisiae (n = 8) in the environments which favour Alpechins’ growth (two-tailed Welsh test, FDR q = 0.05). Box plots are as in Fig. 2. The number of observations (n) results from the combination of the strains in each group and the number of replicates performed in the experiment. Each dot represents the average among the 8 replicates of each of the 16 Alpechins. c, Genotype maps of the Alpechin, Mexican agave, French Guiana and Brazilian bioethanol clades in a shared introgressed region containing the IRC7 (YFR055W) gene. This gene was previously reported to be introgressed in strains belonging to the Wine/European clade72. The IRC7 gene present in the Alpechins has a European S. paradoxus ancestry, while Mexican Agave, Brazilian Bioethanol and French Guiana lineages inherited the S. paradoxus American allele, underlying an independent introgression and maintenance of IRC7 that supports adaptive introgression. Blue and red colours indicate S. cerevisiae and S. paradoxus regions, respectively. Subtelomeric regions are masked genome wide because they lack synteny between S. cerevisiae and S. paradoxus and cannot be aligned for markers genotyping, although inspection of this specific VI-R subtelomeric region in both S. cerevisiae and S. paradoxus genomes indicates that the telomeric repeats are very close to IRC724. The bottom panel shows genetic features annotation of this region from the S. cerevisiae Wine/European strain DBVPG6765.

Source data

Supplementary information

Supplementary Information

This file contains Supplementary Discussions 1-4 and Supplementary References.

Reporting Summary

Supplementary Tables

This file contains Supplementary Tables 1-10.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

D’Angiolo, M., De Chiara, M., Yue, JX. et al. A yeast living ancestor reveals the origin of genomic introgressions. Nature 587, 420–425 (2020). https://doi.org/10.1038/s41586-020-2889-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-2889-1

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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