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 metabolic growth limitations of petite cells lacking the mitochondrial genome

Abstract

Eukaryotic cells can survive the loss of their mitochondrial genome, but consequently suffer from severe growth defects. ‘Petite yeasts’, characterized by mitochondrial genome loss, are instrumental for studying mitochondrial function and physiology. However, the molecular cause of their reduced growth rate remains an open question. Here we show that petite cells suffer from an insufficient capacity to synthesize glutamate, glutamine, leucine and arginine, negatively impacting their growth. Using a combination of molecular genetics and omics approaches, we demonstrate the evolution of fast growth overcomes these amino acid deficiencies, by alleviating a perturbation in mitochondrial iron metabolism and by restoring a defect in the mitochondrial tricarboxylic acid cycle, caused by aconitase inhibition. Our results hence explain the slow growth of mitochondrial genome-deficient cells with a partial auxotrophy in four amino acids that results from distorted iron metabolism and an inhibited tricarboxylic acid cycle.

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: ATP3 variants overcome the slow-growth phenotype of petites.
Fig. 2: Suppressor mutations lead to an increased mitochondrial membrane potential, while morphology and capacity to metabolize glucose remain largely unaffected.
Fig. 3: Amino acid metabolism is perturbed in petites, and their growth defect is rescued by the addition of selected amino acids.
Fig. 4: Aconitase and tricarboxylic acid cycle are inhibited in petites, which is alleviated in evolved petites.
Fig. 5: Iron overload inhibits aconitase, which is required for petite suppression.

Similar content being viewed by others

Data availability

Proteomic data are provided in PRIDE under accession number PXD011715. A list of mass spectrometry files is provided in Supplementary Table 5. Processed metabolomics and proteomics data are provided in Supplementary Data 13. There is no restriction on data availability.

Code availability

Code used for data analysis and creation of figures is available at https://github.com/JakobV/metabolic-growth-limitations-of-petite-cells/.

References

  1. Liu, Z. & Butow, R. A. Mitochondrial retrograde signaling. Annu. Rev. Genet. 40, 159–185 (2006).

    Article  CAS  PubMed  Google Scholar 

  2. Chelstowska, A. & Butow, R. A. RTG genes in yeast that function in communication between mitochondria and the nucleus are also required for expression of genes encoding peroxisomal proteins. J. Biol. Chem. 270, 18141–18146 (1995).

    Article  CAS  PubMed  Google Scholar 

  3. Lill, R. & Mühlenhoff, U. Maturation of iron–sulfur proteins in eukaryotes: mechanisms, connected processes, and diseases. Annu. Rev. Biochem. 77, 669–700 (2008).

    Article  CAS  PubMed  Google Scholar 

  4. Saraste, M. Oxidative phosphorylation at the fin de siècle. Science 283, 1488–1493 (1999).

    Article  CAS  PubMed  Google Scholar 

  5. Eisenberg, T., Büttner, S., Kroemer, G. & Madeo, F. The mitochondrial pathway in yeast apoptosis. Apoptosis 12, 1011–1023 (2007).

    Article  CAS  PubMed  Google Scholar 

  6. Giorgi, C., Marchi, S. & Pinton, P. The machineries, regulation and cellular functions of mitochondrial calcium. Nat. Rev. Mol. Cell Biol. 19, 713–730 (2018).

    Article  CAS  PubMed  Google Scholar 

  7. Foury, F., Roganti, T., Lecrenier, N. & Purnelle, B. The complete sequence of the mitochondrial genome of Saccharomyces cerevisiae. FEBS Lett. 440, 325–331 (1998).

    Article  CAS  PubMed  Google Scholar 

  8. Anderson, S. et al. Sequence and organization of the human mitochondrial genome. Nature 290, 457–465 (1981).

    Article  CAS  PubMed  Google Scholar 

  9. Schatz, G., Haslbrunner, E. & Tuppy, H. Deoxyribonucleic acid associated with yeast mitochondria. Biochem. Biophys. Res. Commun. 15, 127–132 (1964).

    Article  CAS  PubMed  Google Scholar 

  10. Dunn, C. D. & Jensen, R. E. Suppression of a defect in mitochondrial protein import identifies cytosolic proteins required for viability of yeast cells lacking mitochondrial DNA. Genetics 165, 35–45 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Mitchell, P. Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmotic type of mechanism. Nature 191, 144–148 (1961).

    Article  CAS  PubMed  Google Scholar 

  12. Karnkowska, A. et al. A eukaryote without a mitochondrial organelle. Curr. Biol. 26, 1274–1284 (2016).

    Article  CAS  PubMed  Google Scholar 

  13. Tovar, J. et al. Mitochondrial remnant organelles of Giardia function in iron–sulphur protein maturation. Nature 426, 172–176 (2003).

    Article  CAS  PubMed  Google Scholar 

  14. Zubáčová, Z. et al. The mitochondrion-like organelle of Trimastix pyriformis contains the complete glycine cleavage system. PLoS ONE 8, e55417 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Hashiguchi, K. & Zhang-Akiyama, Q.-M. Establishment of human cell lines lacking mitochondrial DNA. Methods Mol. Biol. 554, 383–391 (2009).

    Article  CAS  PubMed  Google Scholar 

  16. Nagley, P. & Linnane, A. W. Mitochondrial DNA deficient petite mutants of yeast. Biochem. Biophys. Res. Commun. 39, 989–996 (1970).

    Article  CAS  PubMed  Google Scholar 

  17. Ephrussi, B., Hottinguer, H. & Tavlitzki, J. Action de l’acriflavine sur les levures. I. La mutation ‘petite colonie’. Ann. Inst. Pasteur 76, 351–367 (1949).

    Google Scholar 

  18. Patananan, A. N., Wu, T.-H., Chiou, P.-Y. & Teitell, M. A. Modifying the mitochondrial genome. Cell Metab. 23, 785–796 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Slonimski, P. La formation des enzymes respiratoires chez la levure (Desoer, 1953).

  20. Ephrussi, B. & Hottinguer, H. Direct demonstration of the mutagenic action of euflavine on baker’s yeast. Nature 166, 956 (1950).

    Article  CAS  PubMed  Google Scholar 

  21. Chen, X. J. & Clark-Walker, G. D. The petite mutation in yeasts: 50 years on. Int. Rev. Cytol. 194, 197–238 (2000).

    Article  CAS  PubMed  Google Scholar 

  22. Mounolou, J.-C. & Lacroute, F. Mitochondrial DNA: an advance in eukaryotic cell biology in the 1960s. Biol. Cell 97, 743–748 (2005).

    Article  CAS  PubMed  Google Scholar 

  23. Corneo, G., Moore, C., Sanadi, D. R., Grossman, L. I. & Marmur, J. Mitochondrial DNA in yeast and some mammalian species. Science 151, 687–689 (1966).

    Article  CAS  PubMed  Google Scholar 

  24. Mounolou, J. C., Jakob, H. & Slonimski, P. P. Mitochondrial DNA from yeast ‘petite’ mutants: specific changes in buoyant density corresponding to different cytoplasmic mutations. Biochem. Biophys. Res. Commun. 24, 218–224 (1966).

    Article  CAS  PubMed  Google Scholar 

  25. Tewari, K. K., Vötsch, W., Mahler, H. R. & Mackler, B. Biochemical correlates of respiratory deficiency. VI. Mitochondrial DNA. J. Mol. Biol. 20, 453–481 (1966).

    Article  CAS  PubMed  Google Scholar 

  26. Rabinowitz, M. & Swift, H. Mitochondrial nucleic acids and their relation to the biogenesis of mitochondria. Physiol. Rev. 50, 376–427 (1970).

    Article  CAS  PubMed  Google Scholar 

  27. Nass, M. M. The circularity of mitochondrial DNA. Proc. Natl Acad. Sci. USA 56, 1215–1222 (1966).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Garipler, G., Mutlu, N., Lack, N. A. & Dunn, C. D. Deletion of conserved protein phosphatases reverses defects associated with mitochondrial DNA damage in Saccharomyces cerevisiae. Proc. Natl Acad. Sci. USA 111, 1473–1478 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Garipler, G. & Dunn, C. D. Defects associated with mitochondrial DNA damage can be mitigated by increased vacuolar pH in Saccharomyces cerevisiae. Genetics https://doi.org/10.1534/genetics.113.149708 (2013).

  30. Veatch, J. R., McMurray, M. A., Nelson, Z. W. & Gottschling, D. E. Mitochondrial dysfunction leads to nuclear genome instability: a link through iron–sulfur clusters. Cell 137, 1247–1258 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Day, M. Yeast petites and small colony variants: for everything there is a season. Adv. Appl. Microbiol. 85, 1–41 (2013).

    Article  PubMed  Google Scholar 

  32. Li, J. et al. Slow growth and increased spontaneous mutation frequency in respiratory-deficient afo1 yeast suppressed by a dominant mutation in ATP3. G3 10, 4637–4648 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Goffeau, A. et al. Life with 6,000 genes. Science 274, 563–567 (1996).

    Google Scholar 

  34. Chen, X. J., Hansbro, P. M. & Clark-Walker, G. D. Suppression of ρ0 lethality by mitochondrial ATP synthase F1 mutations in Kluyveromyces lactis occurs in the absence of F0. Mol. Gen. Genet. 259, 457–467 (1998).

    Article  CAS  PubMed  Google Scholar 

  35. Weber, E. R., Rooks, R. S., Shafer, K. S., Chase, J. W. & Thorsness, P. E. Mutations in the mitochondrial ATP synthase gamma subunit suppress a slow-growth phenotype of yme1 yeast lacking mitochondrial DNA. Genetics 140, 435–442 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Wang, Y., Singh, U. & Mueller, D. M. Mitochondrial genome integrity mutations uncouple the yeast Saccharomyces cerevisiae ATP synthase. J. Biol. Chem. 282, 8228–8236 (2007).

    Article  CAS  PubMed  Google Scholar 

  37. van Leeuwen, J. et al. Exploring genetic suppression interactions on a global scale. Science 354, aag0839 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Puddu, F. et al. Genome architecture and stability in the Saccharomyces cerevisiae knockout collection. Nature 573, 416–420 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Caspeta, L. et al. Biofuels. Altered sterol composition renders yeast thermotolerant. Science 346, 75–78 (2014).

    Article  CAS  PubMed  Google Scholar 

  40. Dean, S., Gould, M. K., Dewar, C. E. & Schnaufer, A. C. Single point mutations in ATP synthase compensate for mitochondrial genome loss in trypanosomes. Proc. Natl Acad. Sci. USA 110, 14741–14746 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Schnaufer, A., Clark-Walker, G. D., Steinberg, A. G. & Stuart, K. The F1–ATP synthase complex in bloodstream stage trypanosomes has an unusual and essential function. EMBO J. 24, 4029–4040 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Goldring, E. S., Grossman, L. I., Krupnick, D., Cryer, D. R. & Marmur, J. The petite mutation in yeast. Loss of mitochondrial deoxyribonucleic acid during induction of petites with ethidium bromide. J. Mol. Biol. 52, 323–335 (1970).

    Article  CAS  PubMed  Google Scholar 

  43. Canelas, A. B. et al. Integrated multilaboratory systems biology reveals differences in protein metabolism between two reference yeast strains. Nat. Commun. 1, 145 (2010).

    Article  PubMed  Google Scholar 

  44. Chen, X. J. & Clark-Walker, G. D. Specific mutations in alpha- and gamma-subunits of F1–ATPase affect mitochondrial genome integrity in the petite-negative yeast Kluyveromyces lactis. EMBO J. 14, 3277–3286 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Caspeta, L. & Nielsen, J. Thermotolerant yeast strains adapted by laboratory evolution show trade-off at ancestral temperatures and preadaptation to other stresses. mBio 6, e00431 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Clark-Walker, G. D., Hansbro, P. M., Gibson, F. & Chen, X. J. Mutant residues suppressing ρ0 lethality in Kluyveromyces lactis occur at contact sites between subunits of F1–ATPase. Biochim. Biophys. Acta 1478, 125–137 (2000).

    Article  CAS  PubMed  Google Scholar 

  47. Kominsky, D. J. & Thorsness, P. E. Expression of the Saccharomyces cerevisiae gene YME1 in the petite-negative yeast Schizosaccharomyces pombe converts it to petite-positive. Genetics 154, 147–154 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kominsky, D. J., Brownson, M. P., Updike, D. L. & Thorsness, P. E. Genetic and biochemical basis for viability of yeast lacking mitochondrial genomes. Genetics 162, 1595–1604 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Vowinckel, J., Hartl, J., Butler, R. & Ralser, M. MitoLoc: a method for the simultaneous quantification of mitochondrial network morphology and membrane potential in single cells. Mitochondrion 24, 77–86 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Rapaport, D., Brunner, M., Neupert, W. & Westermann, B. Fzo1p is a mitochondrial outer membrane protein essential for the biogenesis of functional mitochondria in Saccharomyces cerevisiae. J. Biol. Chem. 273, 20150–20155 (1998).

    Article  CAS  PubMed  Google Scholar 

  51. Schatz, G. Impaired binding of mitochondrial adenosine triphosphatase in the cytoplasmic ‘petite’ mutant of Saccharomyces cerevisiae. J. Biol. Chem. 243, 2192–2199 (1968).

    Article  CAS  PubMed  Google Scholar 

  52. Klingenberg, M. & Rottenberg, H. Relation between the gradient of the ATP/ADP ratio and the membrane potential across the mitochondrial membrane. Eur. J. Biochem. 73, 125–130 (1977).

    Article  CAS  PubMed  Google Scholar 

  53. Giraud, M. F. & Velours, J. The absence of the mitochondrial ATP synthase delta subunit promotes a slow growth phenotype of ρ- yeast cells by a lack of assembly of the catalytic sector F1. Eur. J. Biochem. 245, 813–818 (1997).

    Article  CAS  PubMed  Google Scholar 

  54. Buchet, K. & Godinot, C. Functional F1–ATPase essential in maintaining growth and membrane potential of human mitochondrial DNA-depleted rho degrees cells. J. Biol. Chem. 273, 22983–22989 (1998).

    Article  CAS  PubMed  Google Scholar 

  55. Clark-Walker, G. D. Kinetic properties of F1–ATPase influence the ability of yeasts to grow in anoxia or absence of mtDNA. Mitochondrion 2, 257–265 (2003).

    Article  CAS  PubMed  Google Scholar 

  56. Smith, C. P. & Thorsness, P. E. Formation of an energized inner membrane in mitochondria with a gamma-deficient F1–ATPase. Eukaryot. Cell 4, 2078–2086 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Atkinson, D. E. The energy charge of the adenylate pool as a regulatory parameter. Interaction with feedback modifiers. Biochemistry 7, 4030–4034 (1968).

    Article  CAS  PubMed  Google Scholar 

  58. Merz, S. & Westermann, B. Genome-wide deletion mutant analysis reveals genes required for respiratory growth, mitochondrial genome maintenance and mitochondrial protein synthesis in Saccharomyces cerevisiae. Genome Biol. 10, R95 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Vowinckel, J. et al. Cost-effective generation of precise label-free quantitative proteomes in high-throughput by microLC and data-independent acquisition. Sci. Rep. 8, 4346 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Malecki, M., Kamrad, S., Ralser, M. & Bähler, J. Mitochondrial respiration is required to provide amino acids during fermentative proliferation of fission yeast. EMBO Rep. 21, e50845 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Liu, Z. & Butow, R. A. A transcriptional switch in the expression of yeast tricarboxylic acid cycle genes in response to a reduction or loss of respiratory function. Mol. Cell. Biol. 19, 6720–6728 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Alam, M. T. et al. The metabolic background is a global player in Saccharomyces gene expression epistasis. Nat. Microbiol. 1, 15030 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Mülleder, M. et al. A prototrophic deletion mutant collection for yeast metabolomics and systems biology. Nat. Biotechnol. 30, 1176–1178 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Tsuji, J. et al. The frequencies of amino acids encoded by genomes that utilize standard and nonstandard genetic codes. Bios 81, 22–31 (2010).

    Article  CAS  Google Scholar 

  65. Druseikis, M., Ben-Ari, J. & Covo, S. The Goldilocks effect of respiration on canavanine tolerance in Saccharomyces cerevisiae. Curr. Genet. 65, 1199–1215 (2019).

    Article  CAS  PubMed  Google Scholar 

  66. Regenberg, B., Düring-Olsen, L., Kielland-Brandt, M. C. & Holmberg, S. Substrate specificity and gene expression of the amino acid permeases in Saccharomyces cerevisiae. Curr. Genet. 36, 317–328 (1999).

    Article  CAS  PubMed  Google Scholar 

  67. Ruiz, S. J., van’t Klooster, J. S., Bianchi, F. & Poolman, B. Growth inhibition by amino acids in Saccharomyces cerevisiae. Microorganisms 9, 7 (2021).

    Article  CAS  Google Scholar 

  68. Jacquier, A. Systems biology: supplementation is not sufficient. Nat. Microbiol. 1, 16016 (2016).

    Article  CAS  PubMed  Google Scholar 

  69. Bianchi, F. et al. Asymmetry in inward- and outward-affinity constant of transport explain unidirectional lysine flux in Saccharomyces cerevisiae. Sci. Rep. 6, 31443 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Alam, M. T. et al. The self-inhibitory nature of metabolic networks and its alleviation through compartmentalization. Nat. Commun. 8, 16018 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Nogae, I. & Johnston, M. Isolation and characterization of the ZWF1 gene of Saccharomyces cerevisiae, encoding glucose-6-phosphate dehydrogenase. Gene 96, 161–169 (1990).

    Article  CAS  PubMed  Google Scholar 

  72. Ralser, M. et al. Dynamic rerouting of the carbohydrate flux is key to counteracting oxidative stress. J. Biol. 6, 10 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  73. Campbell, K., Vowinckel, J., Keller, M. A. & Ralser, M. Methionine metabolism alters oxidative stress resistance via the pentose phosphate pathway. Antioxid. Redox Signal. 24, 543–547 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Grüning, N.-M. et al. Pyruvate kinase triggers a metabolic feedback loop that controls redox metabolism in respiring cells. Cell Metab. 14, 415–427 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  75. Celton, M. et al. A comparative transcriptomic, fluxomic and metabolomic analysis of the response of Saccharomyces cerevisiae to increases in NADPH oxidation. BMC Genomics 13, 317 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Kim, D.-H. et al. mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery. Cell 110, 163–175 (2002).

    Article  CAS  PubMed  Google Scholar 

  77. Epstein, C. B. et al. Genome-wide responses to mitochondrial dysfunction. Mol. Biol. Cell 12, 297–308 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Traven, A., Wong, J. M., Xu, D., Sopta, M. & Ingles, C. J. Interorganellar communication. Altered nuclear gene expression profiles in a yeast mitochondrial DNA mutant. J. Biol. Chem. 276, 4020–4027 (2001).

    Article  CAS  PubMed  Google Scholar 

  79. Ruzicka, F. J. & Beinert, H. The soluble ‘high potential’ type iron–sulfur protein from mitochondria is aconitase. J. Biol. Chem. 253, 2514–2517 (1978).

    Article  CAS  PubMed  Google Scholar 

  80. Kispal, G., Csere, P., Prohl, C. & Lill, R. The mitochondrial proteins Atm1p and Nfs1p are essential for biogenesis of cytosolic Fe/S proteins. EMBO J. 18, 3981–3989 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Kaut, A., Lange, H., Diekert, K., Kispal, G. & Lill, R. Isa1p is a component of the mitochondrial machinery for maturation of cellular iron–sulfur proteins and requires conserved cysteine residues for function. J. Biol. Chem. 275, 15955–15961 (2000).

    Article  CAS  PubMed  Google Scholar 

  82. Chen, O. S., Hemenway, S. & Kaplan, J. Inhibition of Fe–S cluster biosynthesis decreases mitochondrial iron export: evidence that Yfh1p affects Fe–S cluster synthesis. Proc. Natl Acad. Sci. USA 99, 12321–12326 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Gillet, L. C. et al. Targeted data extraction of the MS/MS spectra generated by data-independent acquisition: a new concept for consistent and accurate proteome analysis. Mol. Cell. Proteomics 11, O111.016717 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  84. Vowinckel, J. et al. The beauty of being (label)-free: sample preparation methods for SWATH–MS and next-generation targeted proteomics. F1000Res. 2, 272 (2013).

    Article  PubMed  Google Scholar 

  85. Puig, S., Askeland, E. & Thiele, D. J. Coordinated remodeling of cellular metabolism during iron deficiency through targeted mRNA degradation. Cell 120, 99–110 (2005).

    Article  CAS  PubMed  Google Scholar 

  86. Regev-Rudzki, N., Karniely, S., Ben-Haim, N. N. & Pines, O. Yeast aconitase in two locations and two metabolic pathways: seeing small amounts is believing. Mol. Biol. Cell 16, 4163–4171 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Farooq, M. A., Pracheil, T. M., Dong, Z., Xiao, F. & Liu, Z. Mitochondrial DNA instability in cells lacking aconitase correlates with iron citrate toxicity. Oxid. Med. Cell. Longev. 2013, e493536 (2013).

    Article  Google Scholar 

  88. Chen, X. J., Wang, X., Kaufman, B. A. & Butow, R. A. Aconitase couples metabolic regulation to mitochondrial DNA maintenance. Science 307, 714–717 (2005).

    Article  CAS  PubMed  Google Scholar 

  89. Chacinska, A., Koehler, C. M., Milenkovic, D., Lithgow, T. & Pfanner, N. Importing mitochondrial proteins: machineries and mechanisms. Cell 138, 628–644 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Kohlhaw, G. B. Leucine biosynthesis in fungi: entering metabolism through the back door. Microbiol. Mol. Biol. Rev. 67, 1–15 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Ljungdahl, P. O. & Daignan-Fornier, B. Regulation of amino acid, nucleotide, and phosphate metabolism in Saccharomyces cerevisiae. Genetics 190, 885–929 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Tracy, J. W. & Kohlhaw, G. B. Reversible, coenzyme-A-mediated inactivation of biosynthetic condensing enzymes in yeast: a possible regulatory mechanism. Proc. Natl Acad. Sci. USA 72, 1802–1806 (1975).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Irvin, S. D. & Bhattacharjee, J. K. A unique fungal lysine biosynthesis enzyme shares a common ancestor with tricarboxylic acid cycle and leucine biosynthetic enzymes found in diverse organisms. J. Mol. Evol. 46, 401–408 (1998).

    Article  CAS  PubMed  Google Scholar 

  94. Winzeler, E. A. et al. Functional characterization of the S. cerevisiae genome by gene deletion and parallel analysis. Science 285, 901–906 (1999).

    Article  CAS  PubMed  Google Scholar 

  95. Goldstein, A. L. & McCusker, J. H. Three new dominant drug resistance cassettes for gene disruption in Saccharomyces cerevisiae. Yeast 15, 1541–1553 (1999).

    Article  CAS  PubMed  Google Scholar 

  96. Sikorski, R. S. & Hieter, P. A system of shuttle vectors and yeast host strains designed for efficient manipulation of DNA in Saccharomyces cerevisiae. Genetics 122, 19–27 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Klinger, H. et al. Quantitation of (a)symmetric inheritance of functional and of oxidatively damaged mitochondrial aconitase in the cell division of old yeast mother cells. Exp. Gerontol. 45, 533–542 (2010).

    Article  CAS  PubMed  Google Scholar 

  98. Motley, A. M. & Hettema, E. H. Yeast peroxisomes multiply by growth and division. J. Cell Biol. 178, 399–410 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Baganz, F., Hayes, A., Marren, D., Gardner, D. C. & Oliver, S. G. Suitability of replacement markers for functional analysis studies in Saccharomyces cerevisiae. Yeast 13, 1563–1573 (1997).

    Article  CAS  PubMed  Google Scholar 

  100. Schweiger, M. R. et al. Genome-wide massively parallel sequencing of formaldehyde fixed-paraffin embedded tumor tissues for copy-number and mutation analysis. PLoS ONE 4, e5548 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  101. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Wang, K., Li, M. & Hakonarson, H. ANNOVAR: functional annotation of genetic variants from high-throughput sequencing data. Nucleic Acids Res. 38, e164 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  103. Kahm, M., Hasenbrink, G., Lichtenberg-Fraté, H., Ludwig, J. & Kschischo, M. grofit: fitting biological growth curves with R. J. Stat. Softw. 33, 1–21 (2010).

    Article  Google Scholar 

  104. Litsios, A. et al. Differential scaling between G1 protein production and cell size dynamics promotes commitment to the cell division cycle in budding yeast. Nat. Cell Biol. 21, 1382–1392 (2019).

    Article  CAS  PubMed  Google Scholar 

  105. Lamprecht, M. R., Sabatini, D. M. & Carpenter, A. E. CellProfiler: free, versatile software for automated biological image analysis. Biotechniques 42, 71–75 (2007).

    Article  CAS  PubMed  Google Scholar 

  106. Zelezniak, A. et al. Machine learning predicts the yeast metabolome from the quantitative proteome of kinase knockouts. Cell Syst. 7, 269–283 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Mülleder, M. et al. Functional metabolomics describes the yeast biosynthetic regulome. Cell 167, 553–565 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  108. Mülleder, M., Bluemlein, K. & Ralser, M. A high-throughput method for the quantitative determination of free amino acids in Saccharomyces cerevisiae by hydrophilic interaction chromatography–tandem mass spectrometry. Cold Spring Harb. Protoc. 2017, pdb.prot089094 (2017).

    Article  Google Scholar 

  109. Szklarczyk, D. et al. STRING v11: protein–protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 47, D607–D613 (2019).

    Article  CAS  PubMed  Google Scholar 

  110. Jassal, B. et al. The reactome pathway knowledgebase. Nucleic Acids Res. 48, D498–D503 (2020).

    CAS  PubMed  Google Scholar 

  111. MacRae, J. I. et al. Mitochondrial metabolism of sexual and asexual blood stages of the malaria parasite Plasmodium falciparum. BMC Biol. 11, 67 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  112. Behrends, V., Tredwell, G. D. & Bundy, J. G. A software complement to AMDIS for processing GC–MS metabolomic data. Anal. Biochem. 415, 206–208 (2011).

    Article  CAS  PubMed  Google Scholar 

  113. Molik, S., Lill, R. & Mühlenhoff, U. Methods for studying iron metabolism in yeast mitochondria. Methods Cell Biol. 80, 261–280 (2007).

    Article  CAS  PubMed  Google Scholar 

  114. Passonneau, J. V. & Lowry, O. H. in Enzymatic Analysis 229–305 (Humana Press, 1993); https://doi.org/10.1007/978-1-60327-407-4_7

  115. Magri, S., Fracasso, V., Rimoldi, M. & Taroni, F. Preparation of yeast mitochondria and in vitro assay of respiratory chain complex activities. Nat. Protoc. https://doi.org/10.1038/nprot.2010.25 (2010).

  116. Stock, D., Leslie, A. G. & Walker, J. E. Molecular architecture of the rotary motor in ATP synthase. Science 286, 1700–1705 (1999).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank our laboratory members and J. Bähler for critical discussion and comments on the manuscript, and C. Kilian for technical support. This work was supported by the Francis Crick Institute, which receives its core funding from Cancer Research UK (FC001134), the UK Medical Research Council (FC001134) and the Wellcome Trust (FC001134), and received specific funding from the European Research Council (StG 260809 and SYG 951475) and the Wellcome Trust (IA 200829/Z/16/Z), as well as the FWF (Austria) for project P26713 (to M.B.) and a Swiss National Science Foundation Postdoc Mobility fellowship (191052 to J.H.).

Author information

Authors and Affiliations

Authors

Contributions

J.V., J.H. and M. Ralser conceived the project, planned and designed experiments with input from all authors. M. Ralser supervised the project. J.V., J.H. and M. Ralser wrote the manuscript with help from all authors. J.V., J.H., H.M., M.K., K.R., M.A.K., M.M., J.D., M.W., M. Rinnerthaler, J.S.L.Y., S.K.A. and A.L. performed experiments and analysed the data. D.M., B.T., N.Z., C.D.D., J.I.M. and M.B. helped supervise the project and contributed to discussion and interpretation of the results.

Corresponding author

Correspondence to Markus Ralser.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Metabolism thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling Editors: Ashley Castellanos-Jankiewicz; Pooja Jha.

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

Extended data

Extended Data Fig. 1 Growth of wild-type strain is not affected by adaptive laboratory evolution.

Wild-type yeast evolved in chemostats in parallel to corresponding petites that were sampled during the evolution experiment and spotted on solid F1 medium. Colony sizes remain largely unaffected over the course of the experiment. Shown is one of n = 3 biological replicates.

Extended Data Fig. 2 Redox metabolism of wild-type, naive and evolved petites.

a. Wild-type (⍴+) and petite (⍴0) BY4741 or BY4741 Δzwf1 yeast containing empty vector or plasmid-encoded ATP3, ATP3-6 or ATP3-7 were cultured in SC without histidine (SC-H) medium to mid-exponential phase, harvested, and spotted in serial dilutions onto freshly prepared SC-H agar plates ±the oxidant diamide (1.2 mM). Suppressor mutations (ATP3-6, ATP3-7) benefit growth in both BY4741 and BY4741 Δzwf1 background (left panel). Naive petites are slightly more sensitive to diamide compared to evolved petites and wild-type (right panel). In a Δzwf1 background, all strains are hypersensitive to diamide. Shown is one representative experiment n = 3 independent replicates. b.-d. Wild-type (⍴+) and petite (⍴0) YSBN11 strains containing empty vector or plasmid-encoded ATP3, ATP3-6 or ATP3-7 were cultured in SC-H medium to mid-exponential phase and reduced glutathione (GSH) and oxidized glutathione (GSSG) were quantified by LC-MS/MS. In comparison to untreated wild-type controls, treatment with H2O2 (0.5 mM) reduces GSH and increases GSSG levels (positive control). Instead, GSSG and GSH levels are largely unaffected in petites compared to wild-type. Statistics based on two-sided, unpaired t-tests of n = 3 biological replicates comparing wild type to other genotypes. Mean values ± SD. e. Wild-type (⍴+) and petite (⍴0) YSBN11 yeast containing empty vector or plasmid-encoded ATP3-6 were cultured in SM medium, harvested in mid-log phase and re-suspended in medium ± 2 mM tert-BOOH. Superoxide levels were quantified by DHE oxidation after 5 min using fluorescence microscopy. 3 % of wild-type cells, 1 % of naive petite cells and 2 % of evolved petites show DHE fluorescence, while 31 % of wild-type cells challenged with tert-BOOH exhibit fluorescence. Statistics based on two-sided, unpaired t-tests of the following number of biological replicates comparing wild type to other genotypes ⍴+, n = 6 replicates, 13260 cells; ⍴0, n = 11 replicates, 14873 cells; ⍴0 ATP3-6, n = 7 replicates, 17721 cells; ⍴+ test-BOOH, n = 6 replicates, 17721 cells. Mean values ± SEM are shown.

Extended Data Fig. 3 Retrograde response is activated in petites, mitigated in evolved petites, but is not required for suppression of the petite phenotype.

a., b., Wild-type (⍴+) and petite (⍴0) BY4741 strains containing empty vector or plasmid-encoded ATP3-6 and ATP3-7 variants were transformed with a plasmid encoding Pts1-GFP and cultured in SC-HU (histidin, uracil dropout) medium. Cells were harvested in mid-exponential growth phase, fixed with formaldehyde and observed by fluorescence microscopy. Dot-like structures represent peroxisomes (a). Scale bar = 1 μm. Compared to wild-type (10 ± 1) and evolved petites (10 ± 1), naive petites display a significant increase of peroxisomes (17 ± 1) (b). P-values shown based on unpaired, two-sided t-tests, comparing wild type to other genotypes. Mean values ± SEM. Number of cells quantified is indicated by n in the plot. c. Wild-type (⍴+) and petite (⍴0) YSBN11 yeast containing empty vector or plasmid-encoded ATP3-6 were cultured in SM medium. In the mid-exponential growth phase, cells were harvested and subjected to a proteomics workflow. Shown protein fold-changes were normalized to the wild-type. Proteins previously described to be controlled by retrograde response (RTG)1,61,77,78 as well as regulator Rtg2p are significantly upregulated in naive petites (empty plasmid). In comparison, up-regulation in the evolved petite (ATP3-6) is significantly less pronounced. Statistics based on two-sided, unpaired t-tests of n = 3 biological replicates comparing wild type to other genotypes. Mean values ± SD. d. RTG2 or CIT2 are not required for petite suppression. Wild-type (⍴+) BY4741 yeast and strains deleted for RTG2 and CIT2 were transformed with empty plasmid or plasmid encoding for ATP3-6, and mtDNA was depleted. Cells were grown to mid-log growth phase, and equal numbers were spotted in serial dilutions onto SC-H agar. Petites suffer from a growth defect compared to wild-type, which is further amplified in Δrtg2 and Δcit2 strains. Instead, expression of ATP3-6 restores growth irrespective of the genetic background (wild-type, Δrtg2 and Δcit2).

Extended Data Fig. 4 Principal component analysis of proteome changes in response to iron depletion.

Wild-type (⍴+) and petite (⍴0) YSBN11 strains containing empty vector or plasmid-encoded ATP3-6 were cultured in iron-depleted SM medium (- iron) or SM medium containing 200 µg/L iron[III] chloride (+ iron). In the mid-exponential phase, cells were harvested, and proteins were extracted and trypsin-digested. Samples were analyzed by SWATH-MS and protein fold-changes were calculated in comparison to the iron-replete wild-type condition, and data was analyzed using principal component (PC) analysis. In PC1, proteomes were separated according to the petite (slow-growth) phenotype, with evolved petites (expressing ATP3-6) clustering with the wild-type. In PC2, iron-repleted yeast proteomes are separated from iron-depleted proteomes. Combined loadings of all proteins belonging to three GO term groups were plotted (red arrows), showing change of amino acid metabolism along PC1, and up-regulation of iron transport and down-regulation of iron-sulfur proteins (ISPs) along PC2. Data from n = 4 biological replicates per genotype.

Supplementary information

Supplementary Information

Supplementary Figs. 1–22 and references.

Reporting Summary

Supplementary Tables

Supplementary Tables 1–5 and Supplementary Data 1–3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Vowinckel, J., Hartl, J., Marx, H. et al. The metabolic growth limitations of petite cells lacking the mitochondrial genome. Nat Metab 3, 1521–1535 (2021). https://doi.org/10.1038/s42255-021-00477-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s42255-021-00477-6

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research