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:

Mistargeted mitochondrial proteins activate a proteostatic response in the cytosol

Abstract

Most of the mitochondrial proteome originates from nuclear genes and is transported into the mitochondria after synthesis in the cytosol. Complex machineries which maintain the specificity of protein import and sorting include the TIM23 translocase responsible for the transfer of precursor proteins into the matrix, and the mitochondrial intermembrane space import and assembly (MIA) machinery required for the biogenesis of intermembrane space proteins. Dysfunction of mitochondrial protein sorting pathways results in diminishing specific substrate proteins, followed by systemic pathology of the organelle and organismal death1,2,3,4. The cellular responses caused by accumulation of mitochondrial precursor proteins in the cytosol are mainly unknown. Here we present a comprehensive picture of the changes in the cellular transcriptome and proteome in response to a mitochondrial import defect and precursor over-accumulation stress. Pathways were identified that protect the cell against mitochondrial biogenesis defects by inhibiting protein synthesis and by activation of the proteasome, a major machine for cellular protein clearance. Proteasomal activity is modulated in proportion to the quantity of mislocalized mitochondrial precursor proteins in the cytosol. We propose that this type of unfolded protein response activated by mistargeting of proteins (UPRam) is beneficial for the cells. UPRam provides a means for buffering the consequences of physiological slowdown in mitochondrial protein import and for counteracting pathologies that are caused or contributed by mitochondrial dysfunction.

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

Figure 1: The MIA pathway import efficiency regulates protein synthesis and proteasomal activity.
Figure 2: Defects in the presequence import pathway modulate protein synthesis and proteasomal activity.
Figure 3: Mitochondrial precursor proteins stimulate proteasomal activity.
Figure 4: Mistargeted proteins protect cells against stress.

Similar content being viewed by others

Accession codes

Primary accessions

ArrayExpress

Data deposits

RNA-seq data have been submitted to the ArrayExpress database under accession number E-MTAB-3588. The mass spectrometry data have been deposited to the ProteomeXchange Consortium via the PRIDE partner repository with the dataset identifier PXD001495.

References

  1. 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 

  2. Endo, T., Yamano, K. & Kawano, S. Structural insight into the mitochondrial protein import system. Biochim. Biophys. Acta 1808, 955–970 (2011)

    Article  CAS  PubMed  Google Scholar 

  3. Neupert, W. & Herrmann, J. M. Translocation of proteins into mitochondria. Annu. Rev. Biochem. 76, 723–749 (2007)

    Article  CAS  PubMed  Google Scholar 

  4. van der Laan, M., Hutu, D. P. & Rehling, P. On the mechanism of preprotein import by the mitochondrial presequence translocase. Biochim. Biophys. Acta 1803, 732–739 (2010)

    Article  CAS  PubMed  Google Scholar 

  5. Chacinska, A. et al. Essential role of Mia40 in import and assembly of mitochondrial intermembrane space proteins. EMBO J. 23, 3735–3746 (2004)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Stojanovski, D. et al. Mitochondrial protein import: precursor oxidation in a ternary complex with disulfide carrier and sulfhydryl oxidase. J. Cell Biol. 183, 195–202 (2008)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Ong, S. E. et al. Stable isotope labeling by amino acids in cell culture, SILAC, as a simple and accurate approach to expression proteomics. Mol. Cell. Proteomics 1, 376–386 (2002)

    Article  CAS  PubMed  Google Scholar 

  8. Bragoszewski, P., Gornicka, A., Sztolsztener, M. E. & Chacinska, A. The ubiquitin-proteasome system regulates mitochondrial intermembrane space proteins. Mol. Cell. Biol. 33, 2136–2148 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Brodsky, J. L. Cleaning up: ER-associated degradation to the rescue. Cell 151, 1163–1167 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Ciechanover, A. Intracellular protein degradation: from a vague idea through the lysosome and the ubiquitin-proteasome system and onto human diseases and drug targeting. Biochim. Biophys. Acta 1824, 3–13 (2012)

    Article  CAS  PubMed  Google Scholar 

  11. Finley, D. Recognition and processing of ubiquitin-protein conjugates by the proteasome. Annu. Rev. Biochem. 78, 477–513 (2009)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Goldberg, A. L. Protein degradation and protection against misfolded or damaged proteins. Nature 426, 895–899 (2003)

    Article  ADS  CAS  PubMed  Google Scholar 

  13. Bragoszewski, P. et al. Retro-translocation of mitochondrial intermembrane space proteins. Proc. Natl Acad. Sci. USA 112, 7713–7718 (2015)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. Kusmierczyk, A. R., Kunjappu, M. J., Funakoshi, M. & Hochstrasser, M. A multimeric assembly factor controls the formation of alternative 20S proteasomes. Nature Struct. Mol. Biol. 15, 237–244 (2008)

    Article  CAS  Google Scholar 

  15. Le Tallec, B. et al. 20S proteasome assembly is orchestrated by two distinct pairs of chaperones in yeast and in mammals. Mol. Cell 27, 660–674 (2007)

    Article  CAS  PubMed  Google Scholar 

  16. Hirano, Y. et al. Cooperation of multiple chaperones required for the assembly of mammalian 20S proteasomes. Mol. Cell 24, 977–984 (2006)

    Article  CAS  PubMed  Google Scholar 

  17. Ramos, P. C., Hockendorff, J., Johnson, E. S., Varshavsky, A. & Dohmen, R. J. Ump1p is required for proper maturation of the 20S proteasome and becomes its substrate upon completion of the assembly. Cell 92, 489–499 (1998)

    Article  CAS  PubMed  Google Scholar 

  18. Hanssum, A. et al. An inducible chaperone adapts proteasome assembly to stress. Mol. Cell 55, 566–577 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Schubert, U. et al. Rapid degradation of a large fraction of newly synthesized proteins by proteasomes. Nature 404, 770–774 (2000)

    Article  ADS  CAS  PubMed  Google Scholar 

  20. Fischer, M. et al. Protein import and oxidative folding in the mitochondrial intermembrane space of intact mammalian cells. Mol. Biol. Cell 24, 2160–2170 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Frazier, A. E. et al. Pam16 has an essential role in the mitochondrial protein import motor. Nature Struct. Mol. Biol. 11, 226–233 (2004)

    Article  CAS  Google Scholar 

  22. Chacinska, A. et al. Mitochondrial presequence translocase: switching between TOM tethering and motor recruitment involves Tim21 and Tim17. Cell 120, 817–829 (2005)

    Article  CAS  PubMed  Google Scholar 

  23. Bennett, E. J., Bence, N. F., Jayakumar, R. & Kopito, R. R. Global impairment of the ubiquitin–proteasome system by nuclear or cytoplasmic protein aggregates precedes inclusion body formation. Mol. Cell 17, 351–365 (2005)

    Article  CAS  PubMed  Google Scholar 

  24. Escusa-Toret, S., Vonk, W. I. & Frydman, J. Spatial sequestration of misfolded proteins by a dynamic chaperone pathway enhances cellular fitness during stress. Nature Cell Biol. 15, 1231–1243 (2013)

    Article  CAS  PubMed  Google Scholar 

  25. Haynes, C. M. & Ron, D. The mitochondrial UPR — protecting organelle protein homeostasis. J. Cell Sci. 123, 3849–3855 (2010)

    Article  CAS  PubMed  Google Scholar 

  26. Zhao, Q. et al. A mitochondrial specific stress response in mammalian cells. EMBO J. 21, 4411–4419 (2002)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Nargund, A. M., Pellegrino, M. W., Fiorese, C. J., Baker, B. M. & Haynes, C. M. Mitochondrial import efficiency of ATFS-1 regulates mitochondrial UPR activation. Science 337, 587–590 (2012)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Vilchez, D. et al. RPN-6 determines C. elegans longevity under proteotoxic stress conditions. Nature 489, 263–268 (2012)

    Article  ADS  CAS  PubMed  Google Scholar 

  29. Durieux, J., Wolff, S. & Dillin, A. The cell-non-autonomous nature of electron transport chain-mediated longevity. Cell 144, 79–91 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Houtkooper, R. H. et al. Mitonuclear protein imbalance as a conserved longevity mechanism. Nature 497, 451–457 (2013)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  31. Truscott, K. N. et al. A J-protein is an essential subunit of the presequence translocase-associated protein import motor of mitochondria. J. Cell Biol. 163, 707–713 (2003)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Böttinger, L. et al. In vivo evidence for cooperation of Mia40 and Erv1 in the oxidation of mitochondrial proteins. Mol. Biol. Cell 23, 3957–3969 (2012)

    Article  PubMed  PubMed Central  Google Scholar 

  33. Pfanner, N. et al. Uniform nomenclature for the mitochondrial contact site and cristae organizing system. J. Cell Biol. 204, 1083–1086 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Vestweber, D. & Schatz, G. Point mutations destabilizing a precursor protein enhance its post-translational import into mitochondria. EMBO J. 7, 1147–1151 (1988)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Schreiner, B. et al. Role of the AAA protease Yme1 in folding of proteins in the intermembrane space of mitochondria. Mol. Biol. Cell 23, 4335–4346 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nature Methods 9, 357–359 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Lytovchenko, O. et al. The INA complex facilitates assembly of the peripheral stalk of the mitochondrial F1Fo-ATP synthase. EMBO J. 33, 1624–1638 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Wiese, S. et al. Proteomics characterization of mouse kidney peroxisomes by tandem mass spectrometry and protein correlation profiling. Mol. Cell. Proteomics 6, 2045–2057 (2007)

    Article  CAS  PubMed  Google Scholar 

  41. Cox, J. & Mann, M. MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nature Biotechnol. 26, 1367–1372 (2008)

    Article  CAS  Google Scholar 

  42. Cox, J. et al. Andromeda: a peptide search engine integrated into the MaxQuant environment. J. Proteome Res. 10, 1794–1805 (2011)

    Article  ADS  CAS  PubMed  Google Scholar 

  43. Vizcaíno, J. A. et al. ProteomeXchange provides globally coordinated proteomics data submission and dissemination. Nature Biotechnol. 32, 223–226 (2014)

    Article  CAS  Google Scholar 

  44. Vizcaíno, J. A. et al. The PRoteomics IDEntifications (PRIDE) database and associated tools: status in 2013. Nucleic Acids Res. 41, D1063–D1069 (2013)

    Article  PubMed  CAS  Google Scholar 

  45. Maere, S., Heymans, K. & Kuiper, M. BiNGO: a Cytoscape plugin to assess overrepresentation of gene ontology categories in biological networks. Bioinformatics 21, 3448–3449 (2005)

    Article  CAS  PubMed  Google Scholar 

  46. Shannon, P. et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Kruegel, U. et al. Elevated proteasome capacity extends replicative lifespan in Saccharomyces cerevisiae . PLoS Genet. 7, e1002253 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Meisinger, C., Pfanner, N. & Truscott, K. N. Isolation of yeast mitochondria. Methods Mol. Biol. 313, 33–39 (2006)

    CAS  PubMed  Google Scholar 

  49. Tollervey, D. & Mattaj, I. W. Fungal small nuclear ribonucleoproteins share properties with plant and vertebrate U-snRNPs. EMBO J. 6, 469–476 (1987)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Vandesompele, J. et al. Accurate normalization of real-time quantitative RT–PCR data by geometric averaging of multiple internal control genes. Genome Biol. 3, research0034 (2002)

    Article  PubMed  PubMed Central  Google Scholar 

  51. Teste, M. A., Duquenne, M., Francois, J. M. & Parrou, J. L. Validation of reference genes for quantitative expression analysis by real-time RT–PCR in Saccharomyces cerevisiae . BMC Mol. Biol. 10, 99 (2009)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  52. Cankorur-Cetinkaya, A. et al. A novel strategy for selection and validation of reference genes in dynamic multidimensional experimental design in yeast. PLoS ONE 7, e38351 (2012)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kushnirov, V. V. Rapid and reliable protein extraction from yeast. Yeast 16, 857–860 (2000)

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank A. Fergin, B. Knapp, B. Guiard, W. Voos, M. Glickman, A. Gornicka, A. Loniewska-Lwowska, and T. Wegierski for materials, experimental assistance and discussions. Deposition of the data to the ProteomeXchange Consortium is supported by PRIDE Team, EBI. Research in the B.W. laboratory is supported by the Deutsche Forschungsgemeinschaft and the Excellence Initiative of the German Federal & State Governments (EXC 294 BIOSS). Research in the A.C. laboratory was supported by Foundation for Polish Science – Welcome Programme co-financed by the EU within the European Regional Development Fund (L.W., M.E.S. and E.J.), National Science Centre grants 2011/02/B/NZ2/01402 (L.W., U.T. and A.V.) and 2013/11/B/NZ3/00974 (P.C.) and Ministerial Ideas Plus schema 000263 (E.J.). L.W. and U.T. were also supported by National Science Centre grant 2013/08/T/NZ1/00770 and Swiss National Science Foundation postdoctoral fellowship (PP300P3-147899), respectively. P.B. was supported by the National Science Centre grant 2013/11/D/NZ1/02294.

Author information

Authors and Affiliations

Authors

Contributions

P.B. and S.W. are joint second authors. L.W., U.T., P.B., M.E.S., A.V., P.C., S.M. and E.J. performed and analysed biochemical experiments. P.B. and M.L. performed RNA-seq and analyses. S.W. and S.O. performed the mass spectrometric measurements and analyses. A.C., B.W., M.K. and A.D. analysed and supervised the study. A.C and B.W. conceived the project. All authors interpreted the experiments. A.C. wrote the manuscript with the input of other authors.

Corresponding authors

Correspondence to Bettina Warscheid or Agnieszka Chacinska.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Transcriptome and proteome analysis of WT and mia40-4int cells.

a, Distribution of transcripts quantified by RNA-seq in wild-type and mia40-4int cells based on the GO term for translation provided by the Saccharomyces genome database. b, c, Distribution of transcripts and proteins based on the GO term for mitochondrial ribosome provided by the Saccharomyces genome database. d, SILAC-based strategy for the quantitative analysis of alterations in the proteome of yeast cells bearing the mia40-4int mutation. Derivatives of mia40-4int and wild type (mia40-4intS and WTS) were used for SILAC. The mia40-4intS cells grown at permissive temperature 19 °C in heavy medium containing Lys8 and Arg10 were mixed in equal ratio with WTS cells grown in the light medium containing Lys0 and Arg0 for set 1 (two biological replicates) or vice versa for set 2 (one additional biological replicate). e, The mia40-4intS strain and the parental mia40-4int strain exhibited a similar temperature-sensitive phenotype. Both mutants and corresponding wild-type strains were subjected to consecutive tenfold dilutions, spotted on YPD plates and grown at the indicated temperatures. f, The mia40-4intS strain showed no defect in the import of matrix-targeted Su9-DHFR as expected for mia40-4int. The [35S]-labelled precursor of Su9-DHFR was incubated with isolated mitochondria from mia40-4intS and WTS strains for the indicated time points. g, Import of the MIA substrate, Tim9, was decreased in mia40-4intS as expected for mia40-4int. [35S]-labelled precursor of Tim9 was incubated with the isolated mitochondria from mia40-4intS and WTS strains for the indicated time points. f, g, An excess of non-imported precursor was removed by the treatment of mitochondria with proteinase K. Samples were analysed by reducing and non-reducing SDS–PAGE followed by autoradiography. h, Sections of MS survey spectra of SILAC-encoded peptides for the MIA pathway substrates Cox19, Tim12 and Pet191 exhibiting decreased levels in mia40-4intS cells. MS survey spectra were acquired from the experimental set 1 as depicted in Extended Data Fig. 1d. i, Mitochondria were isolated from mia40-4intS and corresponding WTS grown at 19 °C and analysed by western blotting. j, WTS and mia40-4intS yeast were grown to the stationary phase at 19 °C. Cellular extracts for mitochondrial proteins were analysed. The changes in protein abundance in mia40-4intS were as expected for the mia40-4int strain. k, Proteins quantified in three independent biological replicates were plotted according to their P value (log10) against the log10-transformed mia40-4intS /WTS ratios. Proteins with a P value <0.05 and a fold-change in protein abundance >1.5 or <−1.5 were considered upregulated and downregulated and are marked in green and red, respectively. The MIA pathway substrates are highlighted by enlarged circles. WT, wild-type; p, precursor; m, mature; asterisk indicates unspecific band; IAA, iodoacetamide. f, g, i, j, Uncropped blots/gels are in Supplementary Information Fig. 1.

Extended Data Figure 2 Protein abundance in mia40-4int and mia40-4ints .

a, Distribution of proteins quantified by SILAC-MS analysis of WTS and mia40-4intS cells based on the GO terms for cellular components provided by the Saccharomyces genome database. Subcellular localizations are shown for fractions of proteins with a significant 1.5-fold change in abundance and a P value < 0.05 in the mia40-4intS cells. b, GO term enrichment analysis of proteins found to be significantly downregulated (top) or upregulated (bottom) in mia40-4intS. c, WTS and mia40-4intS yeast were grown to the stationary phase at 19 °C. Cellular extracts were analysed for non-mitochondrial proteins. d, WT and mia40-4int yeast were grown in the respiratory medium to the logarithmic phase at 19 °C and shifted for 6 h to 37 °C. Cellular extracts were analysed. No changes in protein abundance in mia40-4int in comparison to wild type were observed. c, d, Uncropped blots are in Supplementary Information Fig. 1. e, f, Yeast were cultured in the full medium with galactose to early logarithmic phase and shifted for 3 h to 37 °C (mia40-4ints ) (e) or 6 h 37 °C (mia40-4int) (f). Cells were treated with 100 µg ml−1 cycloheximide for 10 min. Yeast lysates were fractionated on a 10–50% linear sucrose gradient and absorbance was monitored at 254 nm. The retention of 40S, 60S ribosomal subunits, monosomes (80S) and polysomes is indicated. The monosomes versus polysomes ratio was quantified. Mean ± s.e.m., n = 3. WT, wild type.

Extended Data Figure 3 Characterization of the mia40-4int mutant.

a, Wild type and mia40-4int were grown in respiratory medium and their growth was compared upon shift to the restrictive temperature of 37 °C. D600nm (OD600) was measured at the indicated time points. b, The survival of wild type and mia40-int grown to the logarithmic phase at 19 °C, heat-shocked for 6 h at 37 °C or heat-killed for 3 min at 80 °C was assessed by propidium iodide (PI) staining. c, The mia40-4int and wild-type cells transformed or not transformed with the plasmid encoding Mia40 were grown in respiratory medium at 19 °C and shifted to 37 °C for the indicated times. Protein levels were analysed. The mia40-4int-dependent defect in protein levels was complemented by MIA40. Uncropped blots are in Supplementary Information Fig. 1. d, mRNA levels of IRC25 or POC4 in mia40-4intS, mia40-4int and the corresponding wild type. Cultures were grown in respiratory medium at 19 °C. The heat stress was conducted at 37 °C for the indicated time. Transcript levels of ALG9, FBA1 and TUB2 were used for normalization. Wild type was set to 1. Mean ± s.e.m., n = 3. WT, wild type.

Extended Data Figure 4 Proteasomal activity in mia40 mutants.

a, Wild-type cells were grown in respiratory medium at 19 °C and shifted to 37 °C for 6 h. Where indicated, cell lysates were incubated with 50 µM MG132 for 2 min before addition of caspase-like or chymotrypsin-like proteasome substrate. The activity was inhibited upon MG132 addition, confirming the proteasomal specificity of the assay. b, Wild-type cells and cells deleted for POC1, POC2, IRC25, POC4, UMP1 and PRE9 were grown in respiratory medium to stationary phase at 24 °C. The proteasome caspase-like and chymotrypsin-like activities were reduced upon the compromised proteasome demonstrating the specificity of the assay. Mean ± s.e.m., n = 3. *P value <0.05. c, Cells were grown in respiratory medium to logarithmic phase at 19 °C and shifted to 37 °C for 6 h. Proteasomal activities were analysed over time. Left, mean of 7 biological replicates; right, mean of 4 biological replicates. d, Wild-type cells were grown at 24 °C and were shifted for 6 h at 37 °C or for 4 h at 42 °C. Proteasome activities and the levels of proteasomal subunits, Pre10, Rpt1 and Rpt5 were not changed. Mean ± s.e.m., n = 3. e, Total protein extracts of plasmid-borne mia40 mutants were analysed by anti-ubiquitin immunoblotting and Coomassie staining. d, e, Uncropped blots/gels are in Supplementary Information Fig. 1. f, Proteasomal activity of wild type and mia40 mutants. Mean ± s.e.m., n = 4. *P < 0.05; **P < 0.03. WT, wild type; RFU, relative fluorescent units.

Extended Data Figure 5 Characterization of cells that overproduce Mia40.

a, Formation of a disulfide-bonded intermediate between Mia40 and Tim9 is accelerated in mitochondria with overproduced Mia40. Mitochondria were isolated from wild-type and Mia40 overproducing (Mia40↑) strains and incubated with [35S]-labelled Tim9 precursor. When indicated, iodoacetamide (IAA) was added as a control to block mitochondrial import. b, More efficient import of proteins in mitochondria with overproduced Mia40. Mitochondria from WT and Mia40↑ strains were incubated with [35S]-labelled precursors. The samples were treated with proteinase K to remove non-imported proteins. c, Quantification of [35S]radiolabelled Tim9 and Cox19 import. Mean ± s.e.m., n = 3. d, Cellular and mitochondrial protein levels were analysed in wild-type and Mia40↑ strains by western blotting. The overproduction of Mia40 did not change the protein levels. e, Quantification of Pet191 in mitochondria (M) in wild type and Mia40↑ after 6 h induction (Fig. 1i). f, The WT and Mia40↑ cells producing Pet191Flag were grown in fermentable medium with 2% glucose at 24 °C and shifted to galactose-containing medium for overnight induction. Protein levels in total (T), post-mitochondrial supernatant (S) and mitochondria (M) were analysed. The mitochondrial localization (M) of Pet191 in WT and Mia40↑ after overnight induction was quantified. WT, wild type. a, b, d, f, Uncropped gels/blots are in Supplementary Information Fig. 1.

Extended Data Figure 6 Characterization of mutants that affect the import via the TIM23 presequence pathway.

a, The wild-type strain was grown to stationary phase at 19 °C and treated with CCCP for 2 h to dissipate the electrochemical potential of the inner mitochondrial membrane. The CCCP treatment resulted in the accumulation of precursor proteins. b, Representation of the pulse-chase experiment. The wild type was treated for 30 min with CCCP (pulse) and chased in fresh medium without CCCP for 20 or 45 min for analysis. c, Non-processed precursor proteins accumulate in the tim17 mutants. The tim17 mutants and the corresponding wild type were grown in fermentative medium to stationary phase at 19 °C, shifted to 37 °C and analysed by western blotting. d, The tim17-5 mutant was grown to stationary phase at 19 °C and shifted to 37 °C for 90 min. The cells were fractionated and equal volumes of total (T), post-mitochondrial supernatant (S) and mitochondrial (M) fractions were analysed by western blotting. Precursor proteins were localized in the cytosol together with cytosolic proteins (Rpl17 and Pgk1), in contrast to mature mitochondrial proteins (Cyc3, Tim23, Tom70). e, The pam16 and pam18 mutants were grown to logarithmic growth phase and shifted to restrictive temperature. Total protein content was analysed by western blotting. No changes in protein levels were detected with the exception for a small decrease in the ribosomal proteins Rpl17 and Rpl24 was noticed. f, Ubiquitinated species decreased in the tim17-4 and tim17-5 mutants. The tim17 mutants and corresponding WT strain were grown in respiratory medium at 19 °C, shifted to 37 °C and analysed by anti-ubiquitin immunoblotting. g, The pam16-1 and pam18-1 mutants were grown in respiratory medium at 19 °C and shifted to 37 °C for proteasomal activity assays. Mean of 3 biological replicates. h, The tim17 mutants and corresponding wild-type strain were grown in respiratory medium at 19 °C and shifted to 37 °C for proteasomal activity assays. Mean ± s.e.m., n = 3 (caspase-like activity), n = 5 (chymotrypsin-like activity). ***P < 0.01. i, Northern blot of rRNA and quantification. The pam16-3 and wild-type cells were grown in respiratory medium and shifted to 37 °C. Mean ± s.e.m., n = 3. j, Incorporation of [35S]-labelled amino acids is decreased in tim17 mutants compared to wild type. Strains were grown in respiratory medium and shifted to 37 °C. Samples were taken after 1 or 2 h of [35S] labelling and analysed by SDS–PAGE and autoradiography. k, Representative gradient profiles of ribosomes in pam16-3 and wild type and quantification of the monosome versus polysome fractions. Mean ± s.e.m., n = 3. *P < 0.05. Cells were grown to logarithmic phase, shifted to 37 °C for 3 h and treated with 100 µg ml−1 cycloheximide for 10 min. Lysates were fractionated on a 10–50% linear sucrose gradient and absorbance was monitored at 254 nm. The monosomes versus polysomes ratio was quantified. WT, wild-type. RFU, relative fluorescent units. a, cf, i, j, Uncropped blots/gels are in Supplementary Information Fig. 1.

Extended Data Figure 7 Translation and proteasomal activity in the cells overproducing mitochondrial proteins.

a, Pet191Flag and Mix17Flag were expressed in WT cells at 24 °C. b, Flag-tagged proteins were expressed in WT and when indicated cells were treated with MG132 for 4 h. Fractions of total (T), aggregates (P) and soluble (S) proteins were analysed by western blotting. c, Wild type expressing Pet191Flag or Mix17Flag were grown at 24 °C and analysed for total protein content by western blotting. No changes in protein levels compared to wild type were found, including ribosomal or proteasome subunits. d, Northern blot analysis and quantification of rRNA in cells expressing Mix17Flag. Mean ± s.e.m., n = 3. e, Mistargeted mitochondrial proteins do not alter the rate of translation. Incorporation of [35S]-labelled amino acids in wild type expressing Pet191Flag or Mix17Flag. The expression of Flag-tagged proteins was induced for 6 h at 24 °C. Samples were taken 20 min after initiation of [35S] labelling and analysed by SDS–PAGE and audioradiography. f, Expression of Pet191Flag or Mix17Flag stimulates proteasomal activity. Mean of 3 biological replicates. g, Cox4Flag with (pCox4FLAG) or without mitochondrial presequence (mCox4Flag) was expressed in wild type. h, Mdj1Flag and mMdj1Flag proteins were expressed in WT cells. The presence of Flag-tagged proteins was confirmed by immunoblotting. i, Wild type expressing Mdj1Flag or mMdj1Flag were grown at 24 °C and analysed for total protein levels by western blotting. No changes in protein levels compared to control were found, including ribosomal and proteasome subunits. WT, wild type. RFU, relative fluorescent units. ae, gi, Uncropped blots are in Supplementary Information Fig. 1.

Extended Data Figure 8 Proteasomal activity in the cells overproducing non-mitochondrial proteins.

a, Expression of DHFRFlag and DHFRdsFlag was induced at 24 °C and Pex22Flag at 28 °C in wild type. The presence of Flag-tagged proteins was confirmed by immunoblotting. b, Ubc9ts–GFP was induced in galactose and subsequently wild-type cells were shifted to glucose medium at 37 °C to initiate unfolding. After indicated time points, samples were analysed by western blotting. c, Flag-tagged proteins were expressed in wild type for the indicated time and the proteasome was inhibited with MG132 for 3 h if indicated. Fractions of total (T), aggregates (P) and soluble (S) proteins were analysed and no aggregation was observed. d, e, Proteasomal activity in WT expressing DHFRFlag or DHFRdsFlag grown at 24 °C. Mean ± s.e.m. n = 6 (d). Mean of 6 biological replicates (e). f, Wild-type cells expressing DHFRFlag or DHFRdsFlag were grown at 24 °C and analysed for total protein content. No changes in protein levels were found compared to wild type, including ribosomal and proteasome subunits. g, Proteasomal activity in wild type expressing Pex22Flag grown at 28 °C and the abundance of proteins were not significantly changed. No significant change in proteasomal activity was detected upon expression of Ubc9ts–GFP. Mean ± s.e.m., n = 6. h, The proteasomal stimulation by CCCP and Mix17Flag is additive. Wild-type cells overproducing Mix17Flag were treated with CCCP to measure the chymotrypsin-like activity of the proteasome. In the case of Mix17Flag, proteasomal stimulation was less efficient than stimulation reported for Mix17Flag in Fig. 3a due to change in experimental conditions imposed by the CCCP treatment. Mean ± s.e.m., n = 3. *P < 0.05; **P < 0.03. Analysis of cellular protein content showed no difference in proteasome subunits (Rpt1, Rpt5) or ribosomal protein Rpl17. WT, wild type. RFU, relative fluorescent units. ac, fh, Uncropped blots are in Supplementary Information Fig. 1.

Extended Data Figure 9 Auxiliary factors are required to stimulate proteasome by mitochondrial precursor proteins.

a, Proteasomal activity in wild-type cells expressing simultaneously either Poc1Flag and Poc2Myc or Irc25Flag and Poc4Myc. Mean ± s.e.m., n = 3. *P < 0.05. The overexpression of POC1 and POC2 or IRC25 and POC4 was induced from one plasmid. The overexpression of Irc25Flag and Poc4Myc led to a small increase in the proteasomal activity, in spite of the inability to detect these proteins likely due to tight regulation of their abundance (not shown). b, Proteasomal activity in cells lacking Irc25 or Poc4, expressing Mix17Flag or Pet191FLAG and grown at 24 °C. Mean of 4 biological replicates. c, Affinity purification of the proteasome complex via Pre3TAP from cells grown at 28 °C. Load, 5%; eluate, 100%. Chymotrypsin-like activity of the proteasome bound to the column was measured. The specificity was checked by the treatment of the on-column fraction with proteasomal inhibitor MG132. Activity of the purified proteasome via Pre3TAP was measured upon addition of purified Mix17Flag or Pet191Flag (for 7.5% of the on-column fraction). Mean ± s.e.m., n = 3. Uncropped blots are in Supplementary Information Fig. 1. d, Representation of the proteasome complex affinity purification. The subunits of proteasome are assembled into the 20S catalytic core and the 19S regulatory particle. The core and regulatory particles joined together to form the 26S proteasome. Overexpression of an activator (that is, Mix17) stimulates the 26S proteasome assembly. Thus, upon affinity purification via a TAP-tagged proteasome subunit, more proteasomal subunits representing more assembled proteasomes are found in the eluate in the presence of an activator. WT, wild type. RFU, relative fluorescent units.

Extended Data Figure 10 The proteasome assembly heterodimer Irc25–Poc4 is required to protect cells against stress.

a, Representation of heat stress experiments. Cells overexpressing mitochondrial proteins were exposed to different heat shock conditions and subsequently subjected to lethality assessment (left panel). Cells were exposed to a gradual increase in temperature from 42 °C to 53 °C within 30 min or were incubated at 53 °C for 30 min. Mean lethality values of wild-type cells expressing empty plasmid increased with harsher stress conditions (9% for middle panel; 22% for right panel). The lethality of cells expressing empty plasmid was set to 1. Mean ± s.e.m., n = 5 (middle panel), n = 4 (right panel). **P < 0.03; ***P < 0.01. b, Wild type or cells deleted for the IRC25 or POC4 genes and overproducing Pet191Flag or Cox12Flag protein were cultured on agar plates with sucrose. Consecutive tenfold dilutions of cells were spotted on selective medium plates with either glucose or galactose. Cells were grown at the indicated temperatures. c, Model for cellular responses activated by the mitochondrial protein import and precursor over-accumulation stress. WT, wild type.

Supplementary information

Supplementary Information

This file contains full legends for Supplementary Tables 1-3 and Supplementary images. (PDF 13722 kb)

Supplementary Table 1

This file contains RNA-Seq analysis of the mia40-4int mutant versus wild-type strain – see Supplementary Information file for full legend. (XLSX 1323 kb)

Supplementary Table 2

This file contains SILAC-based proteomics analysis of mia40-4intS mutant versus wild-type - see Supplementary Information file for full legend. (XLSX 974 kb)

Supplementary Table 3

This file contains proteins quantified in SILAC-based proteomics in at least two biological replicates - see Supplementary Information file for full legend. (XLSX 1034 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wrobel, L., Topf, U., Bragoszewski, P. et al. Mistargeted mitochondrial proteins activate a proteostatic response in the cytosol. Nature 524, 485–488 (2015). https://doi.org/10.1038/nature14951

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature14951

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