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.
Similar content being viewed by others
Main
The MIA machinery constitutes one of the essential protein translocation pathways into mitochondria and is responsible for the import of proteins targeted to the intermembrane space (IMS) of mitochondria1,2,3,4. RNA sequencing (RNA-seq) analysis of mia40-4int5,6 (a yeast temperature-sensitive mutant defective for protein import into the IMS) compared with wild-type cells was performed under permissive growth conditions to avoid cumulative consequences of broad mitochondrial dysfunction (Supplementary Table 1). Transcripts for cytosolic ribosomal proteins and proteins involved in translation, but not mitochondrial ribosomal proteins, decreased (Fig. 1a and Extended Data Fig. 1aâc). Thus, changes in the transcriptome could not be directly translated into protein abundance. Next, the cellular proteomes of mia40-4intS and wild-typeS cells were determined under permissive growth conditions using stable isotope labelling by amino acids in cell culture (SILAC)7 (Supplementary Table 2 and Extended Data Fig. 1dâg). Sensitive MIA substrates were significantly decreased in the mia40-4intS proteome (Supplementary Table 3 and Extended Data Fig. 1hâk) because they were not maintained upon mitochondrial import inhibition8.
a, b, Distribution of transcripts and proteins based on the GO term (black dots). c, d, Northern blot analysis of RNA and ratio of 25S/18S rRNAs in yeast grown in respiratory medium. Mean ± s.e.m., n = 3. e, Protein synthesis in cells grown in glycerol. f, Distribution of proteins based on GO terms (black dots). Dashed lines separate proteins of P value <0.05 and fold-change <â1.5 and >1.5. g, Ubiquitination analysis of cellular proteins. h, Proteasomal activity. Mean ± s.e.m., n = 7 (caspase-like activity), n = 4 (chymotrypsin-like activity). i, Cells were fractionated into total (T), post-mitochondrial supernatant (S) and mitochondria (M). Mia40â, Mia40 overproducing strain. j, Proteasomal activity in cells grown at 24 °C. Mean ± s.e.m., n = 4 (caspase-like activity), n = 6 (chymotrypsin-like activity). c, h, j, *P < 0.05; **P < 0.03; ***P < 0.01. WT, wild-type. câe, g, i, Uncropped blots are in Supplementary Information Fig. 1.
We determined the relative abundance of 2,564 proteins. A small fraction (11%) exhibited significant changes in abundance with 112 downregulated and 174 upregulated proteins in mia40-4intS (Supplementary Table 3 and Extended Data Fig. 1k). Localization and Gene Ontology (GO) analysis revealed the modulation of the structure and function of genetic information and reshaping of metabolic networks in mia40-4intS (Extended Data Fig. 2a, b). Consistent with the RNA-seq results (Fig. 1a), a class of proteins that constitute the cytosolic ribosome were downregulated (Fig. 1b and Extended Data Fig. 2b, c). Under restrictive growth conditions an imbalance in the ratio between 25S and 18S rRNA (Fig. 1c, d) and a change in ribosome content with less prominent polysomes in the mia40-4int mutants (Extended Data Fig. 2dâf) were observed. Interestingly, the mia40-4int mutant exhibited a decrease in protein synthesis (Fig. 1e), but extensive cellular death was excluded as a reason for this decrease (Extended Data Fig. 3a, b). The mia40-4int cells transformed with wild-type MIA40 exhibited a normal protein synthesis rate demonstrating that the translation defect depends on mia40-4int (Fig. 1e and Extended Data Fig. 3c).
The ubiquitinâproteasome system (UPS)9,10,11,12 is involved in the degradation of MIA substrate proteins8,13. Abundance of many proteins involved in the UPS did not change in mia40-4intS, including the core proteasomal subunits (Supplementary Table 3 and Extended Data Fig. 2d). Components of the proteasome assembly chaperone complex, Irc25 (also known as Poc3) and Poc4 (refs 14, 15, 16, 17), significantly increased at the protein level in mia40-4intS (Fig. 1f) despite non-significant changes at the mRNA level (Extended Data Fig. 3d). Other assembly chaperones, Ump1, Pba1 (also known as Poc1) and Add66 (also known as Poc2 or Pba2) (refs 14, 15, 16, 17), showed a tendency to increase, whereas the stress-inducible proteasome chaperone Adc17 (also known as Tma17) (ref. 18) did not change (Supplementary Table 3). Under restrictive growth conditions, ubiquitinated protein species decreased in mia40-4int and the effect was complemented by MIA40 (Fig. 1g). The abundance of ubiquitinated species may mirror translation efficiency because newly synthesized proteins constitute a substantial fraction of polyubiquitinated proteins in the cell19. However, the proteasomal inhibitor MG132 attenuated the difference in ubiquitination between wild-type and mutant cells, suggesting acceleration in mia40-4 protein turnover (Fig. 1g). The mia40-4int cells exhibited higher proteasomal activity than wild-type cells and this effect was rescued by MIA40 (Fig. 1h and Extended Data Fig. 4aâc). Applying restrictive temperatures to wild-type cells did not stimulate proteasomal activity (Extended Data Fig. 4d). Furthermore, other plasmid-borne conditional mia40 mutants5,6 with a phenotype milder than mia40-4int showed both a clear decrease in ubiquitinated species and an increase in proteasomal activity (Extended Data Fig. 4e, f).
Overexpression of Mia40 improved protein import into the mitochondrial IMS in human cells20 and in yeast as determined by an in organello assay (Extended Fig. 5aâc) but did not influence the abundance of mitochondrial proteins (Extended Data Fig. 5d). However, when the MIA pathway was challenged by the overproduced Flag fusion protein Pet191Flag8, this protein was more efficiently localized to mitochondria (Fig. 1i and Extended Data Fig. 5e, f). The more productive removal of precursors from the cytosol by mitochondria with elevated Mia40 levels resulted in decreased proteasomal activity (Fig. 1j).
The import of a large fraction of mitochondrial proteins relies on the electrochemical potential of the inner mitochondrial membrane (IM potential)1,2,3,4. Cells treated with the chemical uncoupler carbonyl cyanide m-chlorophenyl hydrazine (CCCP) to dissipate the IM potential accumulated non-imported mitochondrial proteins in their longer unprocessed precursor forms (with the mitochondrial targeting sequence, called presequences1,2,3,4). During the chase in the absence of CCCP, precursor proteins diminished (Fig. 2a and Extended Data Fig. 6a, b). Following the accumulation of mitochondrial precursors in the cytosol, a significant stimulation of the proteasome was observed (Fig. 2b). Similar accumulation of precursor proteins was observed in temperature-sensitive mutants of the TIM23 translocase and motor complex, pam 18, pam16 and tim17 (refs 21, 22) (Fig. 2c and Extended Data Fig. 6c). These precursor forms were found in the cytosol (Fig. 2d and Extended Data Fig. 6d). The pam18 and pam16 mutants did not show any major changes in protein abundance (Extended Data Fig. 6e). However, similar to mia40, the pam18, pam16 and tim17 mutants exhibited a clear decrease in ubiquitinated species and increase in proteasomal activity (Fig. 2e, f and Extended Data Fig. 6fâh). Although we did not detect any change in ribosomal proteins or ribosomal RNA content (Extended Data Fig. 6e, i), protein synthesis was decreased and was accompanied by an increase in monosomes in pam16 cells (Fig. 2g and Extended Data Fig. 6j, k). Thus, the defective import of proteins into mitochondria leads to a response that regulates the cellular protein homeostasis through two arms, the attenuation of the cytosolic synthesis of proteins and activation of the proteasome to prevent the accumulation of mistargeted proteins.
a, b, Cells were treated for 30 min with CCCP (pulse) and chased without CCCP and analysed for protein content and proteasomal activity. Mean ± s.e.m., n = 6. c, The pam18, pam16 and wild-type were analysed for protein content. d, The pam16-3 cells after a 90 min shift to 37 °C were fractionated. Total (T); post-mitochondrial supernatant (S); mitochondria (M); precursor (p); mature (m). e, Cells were analysed for ubiquitination. f, Proteasomal activity. Mean ± s.e.m., n = 3 (caspase-like activity), n = 5 (chymotrypsin-like activity). g, Protein synthesis in cells grown in glycerol. b, f, *P < 0.05; **P < 0.03; ***P < 0.01. WT, wild-type. a, câe, g, Uncropped blots are in Supplementary Information Fig. 1.
To determine whether mistargeted mitochondrial proteins are responsible for proteasomal regulation, we used Pet191Flag and Mix17Flag, which were partly mislocalized to the cytosol when ectopically expressed8 but remained mainly soluble (Fig. 1i and Extended Data Figs 5f and 7a, b). Proteins that formed aggregates were previously shown to inhibit the proteasome23. The steady-state levels of cellular proteins, including ribosomal and proteasomal subunits, rRNA levels and cytosolic translation were unchanged upon the Pet191Flag and Mix17Flag overexpression (Extended Data Fig. 7câe). However, in the presence of mistargeted Pet191Flag or Mix17Flag proteasomal activity increased substantially (Fig. 3a and Extended Data Fig. 7f). Thus, mitochondrial protein mislocalization uncoupled protein degradation and protein synthesis effects.
a, Proteasomal activity upon expression of Pet191Flag or Mix17Flag at 24 °C. Mean ± s.e.m., n = 6. b, Proteasomal activity upon expression of mCox4Flag or Cox4Flag at 28 °C and mMdj1Flag or Mdj1Flag at 24 °C. Mean ± s.e.m., n = 5. c, Proteasomal activity upon Mix17Flag expression and CCCP treatment. Mean ± s.e.m, n = 3. d, Proteasomal activity in cells lacking Irc25 or Poc4 upon expression of Pet191Flag or Mix17Flag at 24 °C. Mean ± s.e.m, n = 3 (caspase-like activity), n = 4 (chymotrypsin-like activity). aâc, *P < 0.05; **P < 0.03; ***P < 0.01. WT, wild-type.
Other mitochondrial proteins, Cox4 and Mdj1 (substrates of the TIM23 pathway) and their non-imported variants that lacked presequences (mCox4 and mMdj1), were overexpressed (Extended Data Fig. 7gâi). The overexpression of Cox4 but not Mdj1, led to a visible accumulation of a cytosolic precursor form (Extended Data Fig. 7g, h). Cox4, mCox4, and mMdj1 (despite its low abundance) stimulated the proteasome, in contrast to fully imported Mdj1 which did not stimulate the proteasome (Fig. 3b). Non-mitochondrial model proteins, including mouse dihydrofolate reductase DHFR and its conformationally destabilized variant DHFRds, temperature sensitive Ubc9tsâGFP24 and native peroxisomal protein Pex22, were overexpressed and remained soluble (Extended Data Fig. 8aâc). They did not significantly affect proteasome activity (Extended Data Fig. 8dâg). To better mimic the stress caused by delayed mitochondrial import, we induced Mix17 in the presence of CCCP. The stimulatory effects of import inhibition and Mix17 overexpression on proteasome activity were additive (Fig. 3c and Extended Data Fig. 8h). Thus, proteins that were delayed in their transport to mitochondriaâbecause of overexpression or the absence of mitochondrial targeting signals or chemical import inhibitionâstimulated proteasomal activity.
No proteasomal activity stimulation by Mix17 was observed in cells lacking Irc25 or Poc4 (Fig. 3d and Extended Data Fig. 9a, b). Indeed, proteasomal subunits were not upregulated (in contrast to Irc25 and Poc4; Fig. 1f and Supplementary Table 3) and the purified proteasome was not stimulated in vitro by Pet191Flag or Mix17Flag (Extended Data Fig. 9c). The 26S proteasome is assembled from components, including the 20S catalytic core and the 19S regulatory particle10,11,12. Scl1TAP (core subunit) and Rpn6TAP (regulatory subunit) were used to purify the 20S core (represented by Pre10) and the 26S proteasome that comprised the core and regulatory particles (represented by Rpt1 and Rpt5) (Fig. 4a, b and Extended Data Fig. 9d). The proteasomal subunits were eluted more efficiently with Scl1TAP and Rpn6TAP upon Mix17 overexpression (Fig. 4c, d). Thus, the stimulatory effect was associated with proteasomal assembly.
a, b, Affinity purification via Scl1TAP or Rpn6TAP. c, d, Affinity purification via Scl1TAP or Rpn6TAP upon 6 h induction of Mix17Flag at 24 °C. Load, 5%; eluate, 100%. The empty vector eluate was set to 1. Mean ± s.e.m, n = 4 (c), n = 5 (d). e, f, Lethality upon overexpression of mitochondrial proteins. Mean ± s.e.m, n = 4 (e), n = 3 (f). Mean lethality values of cells with empty vector: wild type, 19% (e) or wild type, 9.54%; Îirc25, 9.01%; Îpoc4, 7.13% (f), were set to 1. g, Cells were subjected to consecutive tenfold dilutions. h, Lethality of cells. Mean ± s.e.m, n = 3. Mean lethality values of cells with empty plasmid: Îpoc4, 5.33%; Îpoc1, 11.02%; Îpoc2, 11.68%; Îump1, 9.91%; Îpre9, 10.74%, were set to 1. câe, f, h, *P < 0.05; **P < 0.03, ***P < 0.01. WT, wild-type. aâd, Uncropped blots are in Supplementary Information Fig. 1.
Interestingly, heat-shock conditions revealed a gain of stress resistance upon mitochondrial precursor appearance (Fig. 4e, f and Extended Data Fig. 10a). In the cells without Irc25 or Poc4, Mix17Flag and Cox12Flag exerted toxicity under temperature-stress conditions (Fig. 4f). The induction of Pet191Flag, Mix17Flag and Cox12Flag but not DHFRFlag in these cells led to a synthetic growth defect (Fig. 4g and Extended Data Fig. 10b). Cells without Poc1, Poc2, Ump1 or the proteasomal component Pre9 did not exhibit an increase in lethality (Fig. 4h). Thus, the response that was activated by protein mistargeting relied on proteasome assembly governed by the Irc25âPoc4 complex.
The mitochondrial protein import slowdown and precursor over-accumulation stress triggers processes that maintain cellular proteostasis. This includes the inhibition of protein synthesis and activation of the proteasome that originates from mistargeted proteins (Extended Data Fig. 10c). This unfolded protein response activated by protein mistargeting (UPRam) exerts a protective effect on the cells. UPRam differs from UPRmt, another response pathway to compromised mitochondrial function present in higher Eukaryotes, because UPRmt acts by regulating the abundance of chaperones and proteases inside mitochondria25,26,27. Proteasomal activation and mild mitochondrial dysfunction have both been strongly implicated in the positive regulation of longevity28,29,30. The newly identified UPRam causally links defects in mitochondrial biogenesis with proteasomal activity, thus providing a possible explanation for the beneficial effects of mild mitochondrial dysfunction on ageing.
Methods
Data reporting
No statistical methods were used to predetermine sample size. The experiments were not randomized. The investigators were not blinded to allocation during experiments and outcome assessment.
Yeast strains
Yeast Saccharomyces cerevisiae strains are derivatives of YPH499 (MATa, ade2-101, his3-Î200, leu2-Î1, ura3-52, trp1-Î63, lys2-801) or BY4741 (MATa, his3Î1; leu2Î0; met15Î0; ura3Î0). The temperature-sensitive mia40-4int (Fomp2-7int; 305), mia40-4 (YPH-fomp2-7; 176), mia40-3 (YPH-BG-fomp2-8; 178), mia40-F311E (660) and corresponding wild-type (398) strains were described previously5,6,8. The mia40-4int was generated by integration of the plasmid-borne mia40-4 allele into the genome6. This mutant showed a stronger decrease in the accumulation of intermembrane space proteins than mia40-4 (refs 5, 6, 8), but due to a single genomic copy of mia40 was more suitable for SILAC. The SILAC strains are arg8::kanMX4 derivatives of YPH499 (524) and mia40-4int (305) and are referred as to WTS (654) and mia40-4intS (558; a gift from Bernard Guiard). The pam16-1 (YPH-BG-mia1-1; 733), pam16-3 (YPH-BG-mia1-3; 734), pam18-1 (YPH-BG-Mdj3-66; 739), tim17-4 (YPH-BG17-9d; 83), tim17-5 (YPH-BG17-21-7; 85) and corresponding wild-type (738, 524) strains were described previously21,22,31. Strains expressing PRE3 (843) or RPN6 (905) with a TAP tag and deletion strains of PRE9, POC1, POC2, IRC25, POC4, UMP1 in the BY4741 genetic background were purchased from Euroscarf. The strain expressing SCL1 with TAP tag was a gift from Michael Glickman (916). To rescue mia40-4int, the wild-type and mia40-4int yeast cells were transformed with a plasmid containing MIA40 under an endogenous promoter (pGB8220; 85p). To overproduce Mia40, yeast cells were transformed with the TRP1-containing plasmid (pAC8-3, 97p) harbouring MIA40 (YKL195W) under control of the GAL10 promoter. The plasmid for Ubc9tsâGFP overproduction was purchased from Addgene (314p)24. To express PET191 (YJR034W), MIX17 (YMR002W), COX12 (YLR038C), POC1 (YLR199C), POC2 (YKL206C), IRC25 (YLR021W), POC4 (YPL144W), COX4 (YGL187C), MDJ1 (YFL016C), PEX22 (YAL055W) and DHFR encoding mouse dihydrofolate reductase or its destabilized version DHFRds in frame with the FLAG tag under control of the GAL10 promoter, yeast cells were transformed with URA3- or LEU2-containing pAG1 (53p), pAG2 (54p), pEJ1 (334p), pAG3 (55p), pPCh6 (320p), pPCh1 (315p), pMS126 (151p), pMS127 (152p), pMS120 (145p), pMS121 (146p), pUT9 (333p), pMS128 (162p), pMS119 (144p) plasmids. Mix17 (YMR002W) was previously known as Mic17 (refs 8, 32, 33). The mMdj1 and mCox4 constructs correspond to 17â511 and 25â155 amino acid residues of Mdj1 and Cox4, respectively. The sequence for DHFRds contains three destabilizing amino acid substitutions C7S, S42C, N49C (gift from Wolfgang Voos34,35).
Growth conditions
The strains were grown on minimal synthetic medium (0.67% (w/v) yeast nitrogen base, 0.079% (w/v) CSM amino acid mix) containing respiratory carbon source (3% (v/v) glycerol supplemented with 0.05 â 0.2% (v/v) glucose) or fermentable carbon source (2% (v/v) sucrose, 2% (v/v) galactose, 2% (v/v) glucose). To induce the GAL10 promoter for the expression of MIA40, cells were grown on minimal selective medium with 1.5% (v/v) glycerol supplemented with 0.5% (v/v) galactose. To induce the GAL10 promoter for the expression of other proteins, cells were grown on minimal selective medium with 2% (v/v) galactose. In the experiments in which Mia40 and Pet191 were simultaneously overproduced, cells were grown on minimal selective medium with 2% (v/v) galactose. The yeast strains were usually grown at permissive temperature (19 °C or 24 °C) to the early logarithmic growth phase and analysed before or after a shift to a restrictive temperature (37 °C).
RNA-seq
Wild-type and mia40-4int cells were grown at 19 °C in YPG (1% (w/v) yeast extract, 2% (w/v) peptone, 3% (v/v) glycerol) medium to the mid-logarithmic phase (D600nm (OD600) of â¼0.65) in three biological replicates per genotype. RNA was extracted with MagNALyser/MagNA Pure Compact RNA Isolation Kit (Roche) from frozen yeast preserved with RNAlater (Ambion). DNA was digested with TURBO DNA-free Kit (Life Technologies). The RNA quality and concentration were tested with Agilent 2100 Bioanalyzer and Qubit RNA BR Assay (Life Technologies). RNA was spiked with ERCC RNA Spike-In Mix (Life Technologies) and rRNA depleted with Ribo-Zero Gold rRNA Removal Kit (Illumina). RNA-seq libraries were prepared with Ion Total RNA-Seq Kit v2 (Life technologies) with 2 + 10 cycles amplification. Equimolar pool of barcoded libraries was used to prepare sequencing template with Ion PI Template OT2 200 Kit v3 (Life Technologies), followed by sequencing with Ion Proton on Ion PI v2 chip (Life Technologies). The reads were adaptor trimmed and aligned using Torrent Suite and Partek Flow as recommended by Life Technologies. The reads were quality trimmed by Torrent Suite and resulting files were aligned to the Ensembl S. cerevisiae R64 reference genome (GCA_000146045.2) by TopHat2 (ref. 36). Not aligned reads were submitted to the Bowtie2 (ref. 37) alignment with the âvery sensitiveâ setting. Alignments were combined and coordinate sorted with SAMtools38. Differential expression analysis was performed with Partek Genomics Suite (Partek). Transcripts were quantified to the Ensembl R64 GTF annotation file and resulting RPKM values were used for statistical analysis in Partek ANOVA (General Linear Model).
SILAC labelling and LC/MS sample preparation
Yeast strains of WTS and mia40-4intS were grown at 19 °C on respiratory minimal liquid medium (0.67% (w/v) yeast nitrogen base, 0.068% (w/v) CSM amino acid mix lacking arginine and lysine, 3% (v/v) glycerol, 0.02% (w/v) l-proline) supplemented with 18.6 mg lâ1 [13C6/15N4] l-arginine and 19.35 mg lâ1 [13C6/15N2] l-lysine (Sigma-Aldrich) or non-labelled l-arginine and l-lysine. Three separate cultures were grown and differentially labelled using SILAC amino acids including a label-switch. Following mixing of differentially SILAC-labelled WTS and mia40-4intS cells and cell lysis, proteins were acetone-precipitated and analysed using a gel-free and a gel-based approach. For the gel-free approach, proteins were redissolved in 60% MeOH/20 mM NH4HCO3 (pH 7.8) followed by reduction and alkylation of cysteines and tryptic digest as described39. For the gel-based approach, proteins were resuspended in urea buffer (7 M urea/2 M thiourea/30 mM Tris-HCl, pH 8.5) and separated (15 µg protein per replicate) on a 4 â 12% NuPAGE Bis-Tris gel. Lanes were cut into 15 slices, destained and washed as reported40 followed by reduction and alkylation of cysteines39. Proteins were in-gel digested with trypsin and peptides reconstituted in 0.1% TFA.
LC/MS analysis
Peptides obtained by in-solution digestion were analysed in triplicates per biological replicate using an RSLCnano/LTQ-Orbitrap XL system (Thermo Scientific, Bremen, Germany) as described39 with the modification that peptides were separated with a 150-min gradient (4 â 35% acetonitrile in 0.1% formic acid). Peptides derived from the gel were analysed on an RSLCnano/Orbitrap Elite system using a 45-min gradient and parameters as follows: survey scan range, m/z 370â1,700; resolution of 120,000 (m/z 400); top15 method (CID); automatic gain control and maximum fill time, 1 Ã 106 and 200 ms for the orbitrap, 5 Ã 103 and 150 ms for the linear ion trap; normalized collision energy, 35%; activation q, 0.25; activation time, 10 ms; dynamic exclusion, 45 s.
MS data analysis
Data were processed using MaxQuant/Andromeda (version 1.3.0.5)41,42 and searched against the Saccharomyces Genome Database (http://www.yeastgenome.org). LC/MS data from mitochondrial samples were included to support peptide identifications. The database search was performed with default settings including deamination of asparagine and Pro6 as variable modifications. Protein identifications were based on at least one unique peptide. Mass tolerances for precursor and fragment ions were 6 p.p.m. and 0.5 Da, respectively. A false discovery rate of 1% was applied on peptide and protein level. SILAC ratios were calculated based on unique peptides using MaxQuant default settings and âmatch between runsâ. SILAC ratios for proteins were log-transformed and mean log10 ratios across all replicates and P values (two-sided t-test) were determined. Protein groups exhibiting a posterior error probability (PEP) < 0.01, a P value <0.05 and a minimum fold change ±1.5 were classified as up- or downregulated in mia40-4intS cells. Variances of the mean log10 SILAC ratios were consistent across all three biological replicates yielding uniform data. Sample size, data analysis approach and the statistical test chosen fulfill the requirements for quantitative proteome analyses employing similar LC/MS instrumentation and methods. Statistical methods to determine sample size were not applied. The acquired data were normally distributed, thus meeting the assumption of the statistical test. All raw data and original MaxQuant result files have been deposited to the ProteomeXchange Consortium43 via the PRIDE partner repository44 with the data set identifier PXD001495. GO term enrichment analysis of proteins classified as upregulated or downregulated was performed using the BiNGO45 plug-in 2.44 for Cytoscape 2.8 (ref. 46) and the yeast genome as background. Using the built-in BenjaminiâHochberg procedure, raw P values calculated for GO enrichment analysis were corrected for multiple testing. GO terms with a corrected P value <0.05 were considered enriched.
Proteasome activity assays
Proteasome activity assays were carried out as previously described47 with few modifications. 20 D600 nm (OD600) units of cells were shaken in lysis buffer (50 mM Tris-HCl pH 7.4, 10 mM MgCl2, 10% glycerol, 2 mM ATP, 2 mM PMSF, 1 mM DTT) with glass beads (Sigma-Aldrich) for 10â20 min at 4 °C. After a clarifying spin (20,000g, 15 min, 4 °C), protein concentration was determined by Bradford assay with bovine serum albumin as standard. Activity assays were performed in a final volume of 200 µl of lysis buffer with 50 µg of soluble total protein extracts in a 96-well plate by adding 100 µM SucâLeuâLeuâValâTyrâAMC peptide substrate (chymotrypsin-like activity; Bachem, I-1395) or 100 µM AcâNIeâProâNIeâAspâAMC (caspase-like activity; Bachem, I-1850). Fluorescence (excitation wavelength 380 nm, emission wavelength 460 nm) was measured every 5 min for 1â2 h at 25 °C using a microplate fluorometer (Infinite M1000, Tecan). For in vivo inhibition of proteasome activity, cells were grown as described previously8. Cultures were supplemented with either 75 µM MG132 or DMSO.
Stress assays
Cells were grown to the early logarithmic phase and expression of proteins was induced for 6 h at 24 °C in minimal medium (0.67% yeast nitrogen base, 0.077% amino acid supplement mixture) containing 2% (v/v) galactose. Only cultures with similar D600nm (OD600) were compared for their viability after heat shock. One D600nm (OD600) unit of yeast cells was washed with PBS (137 mM NaCl, 12 mM phosphate, 2.7 mM KCl, pH 7.4). Cells were stained with 3 µg mlâ1 propidium iodide in PBS for 15 min at room temperature in dark and analysed within 1 h by flow cytometry (FACSCalibur). 10,000 cells for each sample were measured.
Mitochondria procedures and cell fractionation
Yeast were usually grown on YPG medium. The yeast with overproduced Mia40 were grown on YPGal (1% (w/v) yeast extract, 2% (w/v) peptone, 2% (w/v) galactose). Mitochondria were isolated by differential centrifugation according to the standard method48. Mitochondria were solubilized in Laemmli buffer with 50 mM DTT (reducing) or 50 mM iodoacetamide (non-reducing conditions), and analysed by SDSâPAGE followed by western blotting. [35S]methionine radiolabelled precursors were prepared by in vitro transcription (mMESSAGE mMACHINE, Ambion) and translation (rabbit reticulocyte lysate system, Promega) and imported into the isolated mitochondria as previously described6. Import experiments were performed in at least two different mitochondrial isolations (biological replicates) and repeated at least three times (technical replicates). For fractionation, yeast were grown on minimal medium with fermentable carbon source (2% glucose) to D600nm (OD600) 0.5, induced for 6 h or overnight by 2% galactose for overexpression of MIA40 and 35 D600nm (OD600) units were fractionated by differential centrifugation8.
Mitochondrial precursor protein analysis
Yeast cells were grown in YPGal with addition of 0.2% glucose at 19 °C. Upon consumption of glucose, yeast were grown for additional 2â3 h to D600nm (OD600) 1.5â3. Cultures were then supplemented with either 40 µM CCCP or DMSO. After 1.5 h, samples were analysed by SDSâPAGE and western blotting. For pulse-chase experiments, upon consumption of glucose, yeast were grown for an additional 1 h to D600nm (OD600) 1.1â1.3. Cultures were then supplemented with either 10 µM CCCP or DMSO for the pulse. For pulse-chase experiments with CCCP in combination with expression of Mix17Flag cells were grown at 24 °C in minimal selective medium containing 2% (v/v) sucrose to logarithmic phase. Cells were washed and resuspended in YPGal medium and grown for 4 h at 24 °C. Cells were shifted for 15 min at 19 °C before supplementing the culture with either 10 µM CCCP or DMSO and samples were chased at 19 °C.
Aggregation assay
Cells were grown for 4 h at 28 °C with either 75 µM MG132 or the corresponding volume of DMSO8. Then 10 D600nm (OD600) units of cells were shaken in lysis buffer (30 mM Tris-HCl pH 7.4, 20 mM KCl, 5 mM EDTA, 0.5 mM PMSF, 0.5% Trition X-100) with glass beads for 10 min at 4 °C. After clarifying centrifugation (4,000g for 10 min, 4 °C), total (T) samples were separated into pellet (P) of aggregated proteins and supernatant (S) by ultracentrifugation for 1 h at 125,000g. Total and supernatant fractions were precipitated by 10% trichloroacetic acid.
Affinity purifications
For purification of Flag-tagged proteins, cells were grown at 28 °C in 3% (v/v) glycerol minimal medium with 0.5% (v/v) galactose. Cells (100U D600nm (OD600)) were shaken in buffer A (20 mM Tris-HCl pH 7.4, 300 mM NaCl, 2 mM PMSF) with glass beads for 20 min at 4 °C. The solution was supplemented with 1% (w/v) digitonin and incubated for 15 min on ice. After clarifying centrifugation (20,000g, 15 min, 4 °C), supernatants were incubated with anti-Flag M2 affinity gel (Sigma-Aldrich) at 4 °C for 1 h. The column was washed with buffer A, followed by elution with 0.1 M glycine pH 3.5 for 5 min at room temperature. The pH of the elution fraction was adjusted to pH 7.4.
For proteasome complex purification, 100 D600nm (OD600) units of cells were shaken in buffer B (50 mM Tris-HCl pH 7.4, 10 mM MgCl2, 10% glycerol, 2 mM ATP, 2 mM PMSF, 1 mM DTT) with glass beads for 10 min at 4 °C. After clarifying centrifugation (20,000g, 15 min, 4 °C) supernatants were incubated with immunoglobulin G Sepharose (GE Healthcare) for 1 h at 4 °C. The column was washed and resuspended in 100 µl of buffer B (for purified proteasome activity assay) or proteins were eluted by Laemmli buffer with 50 mM DTT. Proteasome components co-purified with TAP-tagged proteins were quantified and the ratio of signals from cells with Mix17 versus empty plasmid was calculated and normalized to the efficiency of Scl1TAP or Rpn6TAP recovery in the eluate.
Ribosome profile analysis
Yeast were grown on YPGal to early logarithmic phase and shifted to 37 °C. Cells were disrupted in lysis buffer (40 mM Tris-HCl pH 7.4, 100 mM NaCl, 30 mM MgCl2, 100 µg mlâ1 cycloheximide, 300 µg mlâ1 heparin, 1 mM DTT, 0.5 mM PMSF, 2 µM pepstatin A, 2 µg mlâ1 chymostatin, 1 µM leupeptin, 2 mM benzamidie HCl and 5 µM chymostatin) by shaking with glass beads followed by Triton X-100 addition to 0.2% and incubation for 15 min on ice. After clarifying spin (20,000g, 10 min, 4 °C), samples were separated on 10â50% (w/v) sucrose gradients in polysome buffer (20 mM Tris-HCl pH 7.4, 100 mM NaCl, 5 mM MgCl2, 100 µg mlâ1 cycloheximide, 2 mM DTT; 222,000g, 2 h, 4 °C). Gradients were fractionated with a continuous absorbance measurement at 254 nm. The peak areas (volume (ml) à absorbance (mAU)) were quantified with Unicorn 5.2 (GE Healthcare).
Northern blot and qPCR assays
Total RNA was extracted from cells as described49. For northern blot 5 μg of RNA was separated on agarose or acrylamide denaturing gels, transferred to Hybond N+ (GE Healthcare) membrane and immobilized by the UVP CL-1000 crosslinker. Oligonucleotides used for RNA hybridizations (25S rRNA (007) 5â²-CTCCGCTTATTGATATGC; 18S rRNA (008) 5â²-CATGGCTTAATCTTTGAGAC; 5.8S rRNA (017) 5â²-GCGTTGTTCATCGATGC; 5S rRNA 5â²-CTACTCGGTCAGGCTC) were labelled with [γ-32P]ATP using T4 polynucleotide kinase (New England Biolabs). All northern blot steps were performed using PerfectHyb Plus Hybridization Buffer (Sigma-Aldrich). Radioactive signals were detected by digital autoradiography. For qPCR analysis RNA extracts were treated with TURBO DNase (Ambion) and re-isolated. 1 µg of RNA was reverse transcribed using SuperScript III First-Strand Synthesis SuperMix (Invitrogen) with 2.5 µM oligo(dT)20 and 1.8 µM random hexamers as primers. Quantitative PCR reactions were performed using 7900HT Real-Time PCR (LifeTechnologies) and SensiFAST SYBR Hi-ROX Kit (Bioline) with standard protocol in triplicates. Standard curves were generated by the serial dilution of the pooled cDNA. To normalize nonspecific variations, the normalization factor was calculated as the geometric mean of transcript levels of 3 control genes, ALG9, FBA1 and TUB2 that were selected on the basis of stability in transcriptomic and proteomic data sets and published observations50,51,52. Primers used for qPCR: ALG9 5â²-CACGGATAGTGGCTTTGGTGAACAATTAC, 5â²-TATGATTATCTGGCAGCAGGAAAGAACTTGGG; FBA1 5â²-ACGAAGGTCAAAATGCTTCCAT, 5â²-TGGCACAGTGGTCAGAGTGTAAG; IRC25 5â²-CACTGGCAGGAAACCTAGGAAA, 5â²-CTCCCAACTTGGTAACCACTTGA; POC4 5â²-CTCTACCTGCAGACGATCGTTCT, 5â²-GGGAATGGCGTAGTAATAACATGA; TUB2 5â²-ATCTTGTCCCATTCCCACGTT, 5â²-TTGCTGTGTTAATTCAGGGACAGT.
Statistical analysis
For statistical analysis, a two-tailed, paired t-test was used. For statistical analysis of SILAC, polysome profiles, northern blots and qPCR, a two-tailed, unpaired t-test, assuming unequal variance, was performed. A P value <0.05 was considered significant. Proteasomal activity was calculated as a slope value and data are represented in a fold change compared to wild-type strain or wild-type containing the empty plasmid (with the exception for the pulse-chase experiments, in which the CCCP-treated samples were compared to the DMSO-treated control). Represented fold changes are the mean of fold changes obtained from independent biological replicates ± standard error of the mean (s.e.m.). ImageQuant was used for the quantification of western blotting results. The sample size used was estimated based on previous experiments and statistical tests were chosen based on published data with comparable methodology. Experiments have been replicated at least three times.
Other details
Volcano plot distribution of transcripts and proteins was based on the GO Slim terms provided by the Saccharomyces genome database (SGD). Protein extracts were prepared from amount of cells corresponding to 0.2â0.4 D600nm (OD600) units by alkaline lysis53 and resuspended in Laemmli buffer containing 50 mM DTT. To assess protein synthesis, cells were grown in Met-free minimal medium before incubating with [35S]-labelled methionine and cysteine (EXPRESS; Perkin Elmer) that were added to a final concentration of 6 µCi mlâ1 for 20 min to 2 h. The samples were analysed by SDSâPAGE and digital autoradiography. Proteins were separated by SDSâPAGE on 15%, 12% or 10% gels. Primary antibodies were custom-raised in rabbits and individually controlled for specificity. Commercially available antibodies for mouse monoclonal anti-ubiquitin (SC-8017, Santa Cruz), anti-Flag M2 (F1804, Sigma-Aldrich), rabbit polyclonal anti-Rpt5 (PW8245, Enzo Life Science), anti-Rpt1 (PW8255, Enzo Life Science), anti-alpha7/Pre10 (PW8110, Enzo Life Science) and anti-Hsp104 (ADI SPA 1040, Enzo Life Science) were used. Chemiluminescence protein signals were detected by use of X-ray films. Autoradiography signals were processed with ImageQuant (GE Healthcare). The images were processed digitally using Adobe Photoshop CS4. In some figures, non-relevant gel parts were excised digitally. Protein concentration was determined using Roti-Quant (Carl Roth GmbH) with bovine serum albumin as a standard. The nomenclature of proteins is according to the SGD.
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
Chacinska, A., Koehler, C. M., Milenkovic, D., Lithgow, T. & Pfanner, N. Importing mitochondrial proteins: machineries and mechanisms. Cell 138, 628â644 (2009)
Endo, T., Yamano, K. & Kawano, S. Structural insight into the mitochondrial protein import system. Biochim. Biophys. Acta 1808, 955â970 (2011)
Neupert, W. & Herrmann, J. M. Translocation of proteins into mitochondria. Annu. Rev. Biochem. 76, 723â749 (2007)
van der Laan, M., Hutu, D. P. & Rehling, P. On the mechanism of preprotein import by the mitochondrial presequence translocase. Biochim. Biophys. Acta 1803, 732â739 (2010)
Chacinska, A. et al. Essential role of Mia40 in import and assembly of mitochondrial intermembrane space proteins. EMBO J. 23, 3735â3746 (2004)
Stojanovski, D. et al. Mitochondrial protein import: precursor oxidation in a ternary complex with disulfide carrier and sulfhydryl oxidase. J. Cell Biol. 183, 195â202 (2008)
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, 376â386 (2002)
Bragoszewski, P., Gornicka, A., Sztolsztener, M. E. & Chacinska, A. The ubiquitin-proteasome system regulates mitochondrial intermembrane space proteins. Mol. Cell. Biol. 33, 2136â2148 (2013)
Brodsky, J. L. Cleaning up: ER-associated degradation to the rescue. Cell 151, 1163â1167 (2012)
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, 3â13 (2012)
Finley, D. Recognition and processing of ubiquitin-protein conjugates by the proteasome. Annu. Rev. Biochem. 78, 477â513 (2009)
Goldberg, A. L. Protein degradation and protection against misfolded or damaged proteins. Nature 426, 895â899 (2003)
Bragoszewski, P. et al. Retro-translocation of mitochondrial intermembrane space proteins. Proc. Natl Acad. Sci. USA 112, 7713â7718 (2015)
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, 237â244 (2008)
Le Tallec, B. et al. 20S proteasome assembly is orchestrated by two distinct pairs of chaperones in yeast and in mammals. Mol. Cell 27, 660â674 (2007)
Hirano, Y. et al. Cooperation of multiple chaperones required for the assembly of mammalian 20S proteasomes. Mol. Cell 24, 977â984 (2006)
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, 489â499 (1998)
Hanssum, A. et al. An inducible chaperone adapts proteasome assembly to stress. Mol. Cell 55, 566â577 (2014)
Schubert, U. et al. Rapid degradation of a large fraction of newly synthesized proteins by proteasomes. Nature 404, 770â774 (2000)
Fischer, M. et al. Protein import and oxidative folding in the mitochondrial intermembrane space of intact mammalian cells. Mol. Biol. Cell 24, 2160â2170 (2013)
Frazier, A. E. et al. Pam16 has an essential role in the mitochondrial protein import motor. Nature Struct. Mol. Biol. 11, 226â233 (2004)
Chacinska, A. et al. Mitochondrial presequence translocase: switching between TOM tethering and motor recruitment involves Tim21 and Tim17. Cell 120, 817â829 (2005)
Bennett, E. J., Bence, N. F., Jayakumar, R. & Kopito, R. R. Global impairment of the ubiquitinâproteasome system by nuclear or cytoplasmic protein aggregates precedes inclusion body formation. Mol. Cell 17, 351â365 (2005)
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, 1231â1243 (2013)
Haynes, C. M. & Ron, D. The mitochondrial UPR â protecting organelle protein homeostasis. J. Cell Sci. 123, 3849â3855 (2010)
Zhao, Q. et al. A mitochondrial specific stress response in mammalian cells. EMBO J. 21, 4411â4419 (2002)
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, 587â590 (2012)
Vilchez, D. et al. RPN-6 determines C. elegans longevity under proteotoxic stress conditions. Nature 489, 263â268 (2012)
Durieux, J., Wolff, S. & Dillin, A. The cell-non-autonomous nature of electron transport chain-mediated longevity. Cell 144, 79â91 (2011)
Houtkooper, R. H. et al. Mitonuclear protein imbalance as a conserved longevity mechanism. Nature 497, 451â457 (2013)
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, 707â713 (2003)
Böttinger, L. et al. In vivo evidence for cooperation of Mia40 and Erv1 in the oxidation of mitochondrial proteins. Mol. Biol. Cell 23, 3957â3969 (2012)
Pfanner, N. et al. Uniform nomenclature for the mitochondrial contact site and cristae organizing system. J. Cell Biol. 204, 1083â1086 (2014)
Vestweber, D. & Schatz, G. Point mutations destabilizing a precursor protein enhance its post-translational import into mitochondria. EMBO J. 7, 1147â1151 (1988)
Schreiner, B. et al. Role of the AAA protease Yme1 in folding of proteins in the intermembrane space of mitochondria. Mol. Biol. Cell 23, 4335â4346 (2012)
Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013)
Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nature Methods 9, 357â359 (2012)
Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078â2079 (2009)
Lytovchenko, O. et al. The INA complex facilitates assembly of the peripheral stalk of the mitochondrial F1Fo-ATP synthase. EMBO J. 33, 1624â1638 (2014)
Wiese, S. et al. Proteomics characterization of mouse kidney peroxisomes by tandem mass spectrometry and protein correlation profiling. Mol. Cell. Proteomics 6, 2045â2057 (2007)
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, 1367â1372 (2008)
Cox, J. et al. Andromeda: a peptide search engine integrated into the MaxQuant environment. J. Proteome Res. 10, 1794â1805 (2011)
VizcaÃno, J. A. et al. ProteomeXchange provides globally coordinated proteomics data submission and dissemination. Nature Biotechnol. 32, 223â226 (2014)
VizcaÃno, J. A. et al. The PRoteomics IDEntifications (PRIDE) database and associated tools: status in 2013. Nucleic Acids Res. 41, D1063âD1069 (2013)
Maere, S., Heymans, K. & Kuiper, M. BiNGO: a Cytoscape plugin to assess overrepresentation of gene ontology categories in biological networks. Bioinformatics 21, 3448â3449 (2005)
Shannon, P. et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498â2504 (2003)
Kruegel, U. et al. Elevated proteasome capacity extends replicative lifespan in Saccharomyces cerevisiae . PLoS Genet. 7, e1002253 (2011)
Meisinger, C., Pfanner, N. & Truscott, K. N. Isolation of yeast mitochondria. Methods Mol. Biol. 313, 33â39 (2006)
Tollervey, D. & Mattaj, I. W. Fungal small nuclear ribonucleoproteins share properties with plant and vertebrate U-snRNPs. EMBO J. 6, 469â476 (1987)
Vandesompele, J. et al. Accurate normalization of real-time quantitative RTâPCR data by geometric averaging of multiple internal control genes. Genome Biol. 3, research0034 (2002)
Teste, M. A., Duquenne, M., Francois, J. M. & Parrou, J. L. Validation of reference genes for quantitative expression analysis by real-time RTâPCR in Saccharomyces cerevisiae . BMC Mol. Biol. 10, 99 (2009)
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)
Kushnirov, V. V. Rapid and reliable protein extraction from yeast. Yeast 16, 857â860 (2000)
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 â 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
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
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 SDSâPAGE 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 <â1.5 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 mlâ1 cycloheximide for 10 min. Yeast lysates were fractionated on a 10â50% 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 (Mia40â) 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 Mia40â 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 Mia40â 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 Mia40â after 6 h induction (Fig. 1i). f, The WT and Mia40â 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 Mia40â 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 SDSâPAGE 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 mlâ1 cycloheximide for 10 min. Lysates were fractionated on a 10â50% 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, câf, 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 SDSâPAGE 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. aâe, gâi, 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, Ubc9tsâGFP 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 Ubc9tsâGFP. 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. aâc, fâh, 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 Irc25âPoc4 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 â 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)
Rights and permissions
About this article
Cite this article
Wrobel, L., Topf, U., Bragoszewski, P. et al. Mistargeted mitochondrial proteins activate a proteostatic response in the cytosol. Nature 524, 485â488 (2015). https://doi.org/10.1038/nature14951
Received:
Accepted:
Published:
Issue date:
DOI: https://doi.org/10.1038/nature14951
This article is cited by
-
Rescue of mitochondrial import failure by intercellular organellar transfer
Nature Communications (2024)
-
Prefoldin 2 contributes to mitochondrial morphology and function
BMC Biology (2023)
-
A cytosolic surveillance mechanism activates the mitochondrial UPR
Nature (2023)
-
Immunoproteasome-specific subunit PSMB9 induction is required to regulate cellular proteostasis upon mitochondrial dysfunction
Nature Communications (2023)
-
The interactome of CLUH reveals its association to SPAG5 and its co-translational proximity to mitochondrial proteins
BMC Biology (2022)






