Abstract Eukaryotic cells can survive the loss of their mitochondrial genome, but consequently suffer from severe growth defects. Since the discovery of this phenotype in the 1940s, ‘petite yeast’ remained instrumental for studying mitochondrial function and physiology. However, the molecular cause of the slow-growth defect remained an open question. Here, we used adaptive laboratory evolution, and identified recurrent mutations within the ATP synthase gene ATP3 that restore the growth rate of petite Saccharomyces cerevisiae cells. Then, we use a combination of molecular genetics, and -omics approaches to pinpoint the metabolic limitations overcome by fast growing petites. Changes in ATP metabolism, mitochondrial morphology or retrograde signalling were detected in petites, but did not explain the growth recovery in evolved petites. Instead, we found that the synthesis of glutamate, glutamine, leucine and arginine is growth limiting in petites. The evolution of fast growth did overcome these amino acid deficiencies, by alleviating a defect in mitochondrial iron metabolism and by restoring a previously overlooked defect in the mitochondrial TCA cycle, caused by aconitase inhibition. Our results hence explain the slow growth of mitochondrial genome-deficient cells with a partial auxotrophy in four amino acids that results from distorted iron metabolism and an inhibited TCA cycle. Introduction Mitochondria are key organelles in eukaryotic cells, and function in a range of crucial physiological processes, from metabolism, calcium and iron homeostasis, signalling, to the execution of apoptosis ^[62]1–[63]6 . In most eukaryotes, mitochondria maintain their own genome, a remnant of their likely α-proteobacterial origin ^[64]7–[65]9 . The majority of genes encoding mitochondrial proteins have been transferred to the nucleus, but there remain important proteins encoded by mitochondrial DNA (mtDNA)^ [66]7,[67]8 . These are translated directly in the mitochondrion and are required for mitochondrial genome function, oxidative phosphorylation, and in turn, the TCA cycle ^[68]1,[69]7,[70]8 . Further, the respiratory chain establishes a chemiosmotic proton gradient at the inner mitochondrial membrane, which drives mitochondrial transport processes, and oxidative phosphorylation ^[71]10,[72]11 . Some organisms with a mostly parasitic lifestyle have lost the mitochondrial genome and evolved without it ^[73]12–[74]14 . Indeed, mtDNA is not under all circumstances essential for basic eukaryotic cell functionality. Spontaneous mitochondrial genome loss can be observed across multiple cell types and species, including human cell lines and yeasts ^[75]15,[76]16 . However, cells that tolerate the spontaneous loss of the mtDNA grow slowly. In yeast, this phenotype has been described more than 70 years ago, and named the “petite” phenotype, as these slow growing yeast cells form small colonies ^[77]17 . Petites are commonly referred to as ρ^0 (opposed to wild type ρ^+ cells), and to our knowledge, constitute the earliest report of a defect directly associated with mitochondria. Accordingly, petites became a frequently used tool to study mitochondrial function. Some discoveries made with petite cells were of historical importance. For example, the petite phenotype led to the discovery of the mitochondrial genome itself ^[78]9 . As the research on mitochondrial function is still hindered by difficulties in the manipulation of the sequence of mtDNA ^[79]18 , petites remain an important model to map the associated spectrum of metabolic and physiological roles of mitochondria, such as ATP production through oxidative phosphorylation ^[80]9,[81]17,[82]19–[83]32 . The growth defect of petites can be compensated by secondary suppressor mutations. For instance, petite suppressors are common in the laboratory S. cerevisiae strain S288C ^[84]33 which spontaneously loses mtDNA, a property known as a high petite frequency ^[85]34 . Petite suppressor mutations frequently map to the F1 subunit of the ATPase (complex V) of the respiratory chain ^[86]21,[87]32,[88]35–[89]39 . Other mutations that affect the F1 subunit of the ATPase have been associated with thermotolerance ^[90]40 , or enable species such as K. lactis and T. brucei where mtDNA is otherwise essential, to survive in its absence ^[91]35,[92]41,[93]42 . Despite knowing that the petite growth establishes as a direct consequence of a lack of the mitochondrial genome, and that the defect can be overcome by mutations in the nuclear encoded subunits of the ATP synthase, it remained unclear which metabolic constraints are causing the slow growth phenotype. We speculated that elucidating the molecular changes observed in evolved petites would ultimately guide us to the underlying problem that restricts growth in the mitochondrially deficient cells. We first conducted an adaptive evolution experiment to select for S. cerevisiae petite mutants that effectively suppress their slow-growth phenotype. Petites adapted to near wild type growth rates despite the absence of mtDNA, by obtaining a set of rapidly and independently occurring, dominant mutations in ATP3 gene, that are close proximity to previously described alleles ^[94]21,[95]35–[96]39 . We then used the newly isolated suppressors as a tool to identify the metabolic bottlenecks that slow growth. Many cellular processes associated with the function of mitochondria, including redox metabolism, nutrient- or retrograde-signalling, were found to be altered in petites. However, they were not causative to their slow-growth phenotype. Instead, we find that petites suffer from an insufficient capacity in the biosynthesis of the four amino acids, glutamate, glutamine, leucine, and arginine. We show that petite cells overcome this multiple, partial auxotrophy by restoring defects in mitochondrial iron metabolism and a so far overlooked block in the TCA cycle at the step of aconitase. Results Adaptive evolution selects for a specific hotspot of mutations in ATP3 that overcome the petite phenotype To test which mutations best suppress the petite phenotype, we designed an adaptive evolution experiment. Using the method by Goldring et al ^[97]43 , we generated mtDNA-deficient (ρ^0) petite progenitors from a diploid, prototrophic Saccharomyces cerevisiae strain (YSBN1) ^[98]44 . We selected clones that formed characteristic petite colonies, and confirmed loss of mtDNA ([99]Fig. 1a, b). Petite and wild type control cultures were then grown in a chemostat under glucose-limiting conditions in a highly nutrient restricted minimal (F1) medium ^[100]44 . The cultures were kept under static nutritional conditions for 65 days, during which approximately 352 biomass doublings occurred ([101]Suppl. Fig. 1). Samples were extracted at regular intervals over time, and were genotyped and phenotyped. Large colonies indicative of improved growth in petite cultures were readily apparent at the second sampling point (65 biomass doublings) and became more abundant with time. After 162 doublings, large colonies dominated all three parallel chemostats ([102]Fig. 1b). Growth of wild type cells cultivated as controls in parallel chemostats remained unaffected ([103]Suppl. Fig. 2). Figure 1. Adaptive laboratory evolution selects for ATP3 variants that overcome the slow-growth phenotype of petites. [104]Figure 1 [105]Open in a new tab A. Petites lost their mitochondrial genome. Wild type (ρ^+) yeasts (YSBN1) and respective progenitor strains that lost their mitochondrial genome following ethidium bromide treatment were analyzed using DAPI staining (3 replicates each, c1 - c3). While wild type yeast contains dot-like structures indicative of mtDNA, petite (ρ^0) yeast only display nuclear DNA staining. Representative images are shown. Scale bar = 1 μm. B. Evolved petites with improved growth emerge during adaptive laboratory evolution. Petite YSBN11 strains (ρ^0) were grown in F1-media in a chemostat under glucose-limiting conditions. Samples were taken from the inoculum, after 65, 162, as well as 260 generations, and spotted on F1 agar plates. After 260 generations, almost all cultures had regained the capacity to form larger colonies. Shown is one representative of n = 3 biological replicates. C. ATP3 substitutions emerge and their prevalence increases during adaptive laboratory evolution. Relative frequency of the three most abundant heterozygous ATP3mutations (ATP3-6, ATP3-7 and ATP3-8) in chemostat cultures was estimated from whole-population sequencing. Prevalence of these SNPs increased during the evolution experiment, and were recovered from 79.2 % of the isolates from the selected evolved population. D. Molecular context of isolated suppressor mutations in the yeast F1c10 ATPase complex. Detailed view of F1c10 ATPase complex structure with α (green), β (turquoise) and γ (grey) subunit interface. Identified and confirmed mutations (shown in [106]Suppl. Fig. 3) in the γ subunit are highlighted in red. The small inlet structure shows the entire yeast F1c10 ATPase complex (PDB ID 2XOK ^[107]118 ) with highlighted γ subunit (dark grey) and mutations in C-terminal region (red). E. Petite isolates from evolved populations that carry ATP3 mutations display elevated growth rates. Fast-growing colonies from evolved populations were genotyped for ATP3, and growth rates were recorded in F1 medium with 0.5 % glucose. The petite ancestor strain exhibits a significantly slower growth rate, compared to colonies carrying variants of ATP3-6, ATP3-7 or ATP3-8. Statistics based on unpaired t-tests comparing naive petite (ρ^0) with other indicated genotypes. *p < 0.05; Shown are mean values ± SD of n > 4 biological replicates per genotype. F, G. Heterologous expression of identified ATP3 substitutions increases growth in naive petites. YSBN11 strains were transformed with pRS313 plasmids carrying indicated ATP3 variants, and mtDNA was depleted. Growth in F1 medium containing 0.5 % glucose was monitored. (G) Expression of ATP3 variants significantly improves growth compared to vector controls, or additional expression of the wild type allele. (F) Petite (ρ^0) cells expressing ATP3 variants display growth advantages. Statistics based on two-sided, unpaired t-tests compared with other genotypes. *p < 0.05. Mean values ± SD of n = 3 biological replicates per genotype are shown. To identify the genetic alterations that had occurred, we performed whole-genome resequencing at the population level ([108]Suppl. Fig. 1, sequenced at generations 0, 260, 340 for all replicates, and additionally generation 65 and 162 for replicate 3). In order to rule out secondary (nuclear) mutations caused by mtDNA depletion, we additionally sequenced respective progenitor strains ([109]Suppl. Table 1). Minority allele detection analysis revealed a surprisingly specific set of 21 non-synonymous and 30 synonymous mutations which were present in the evolved petites, but not in the ancestral, naive petites, nor in the wild type parent ([110]Suppl. Table 1). Mutations only within one gene (ATP3) enriched over time. Moreover ATP3 mutations were detected in each of the independent evolution experiments ([111]Fig. 1c, [112]Suppl. Table 1). We then randomly selected fast-growing colonies of one evolved population (replicate 3), and re-sequenced the ATP3 locus. This analysis confirmed the three major ATP3 variants as detected by the population sequencing, and revealed four additional ATP3 mutations with lower prevalence ([113]Suppl. Fig. 3). Each clone contained exclusively one of the identified ATP3 mutations ([114]Suppl. Fig. 3), indicating that even within one chemostat culture, multiple ATP3 mutations emerged independently. At the final time-point the majority of cells contained ATP3 mutations ([115]Fig. 1c, [116]Suppl. Fig. 3, [117]4). Most uncovered mutations were located at the C-terminal domain of ATP3, and were closely located to previously isolated petite suppressors ^[118]32,[119]36,[120]42,[121]45–[122]47 . This included two mutations that did match previously identified residues associated with slow-growth suppression in a petite S. cerevisiaeyme1 mutant ([123]Suppl. Fig. 4a,[124]b) ^[125]36 . All variants affected conserved residues. Mapping the identified mutations into the structure of complex V highlights a hotspot in a narrow region close to the C-terminus of the ɣ subunit ([126]Fig. 1d, [127]Suppl. Fig. 4). Next, we tested whether the emergence of the ATP3 alleles was causal for the improved growth. We measured growth rate in isolates harbouring ATP3 mutations. Their growth rate was significantly increased compared to the ancestral petite cultures ([128]Fig. 1e). Then, we cloned the two most frequent alleles, ATP3-6 and ATP3-7 ([129]Suppl. Fig. 4a) and reintroduced them on a plasmid into the parental wild type background and naïve (non-evolved) petite cells. In parallel, we constructed an artificial variant which contained the two most frequent mutations in a single gene (named ATP3-6,7). The expression of mutant alleles ATP3-6, ATP3-7, as well as the artificial allele carrying both mutations (ATP3-6,7) restored growth of petites, suggesting a dominant mode of inheritance in petites ([130]Fig. 1f, g). In contrast, the plasmid control and an additional copy of ATP3 did not compensate for the growth defect petites ([131]Fig. 1f, g) nor did their expression affect wild type cells ([132]Suppl. Fig. 5). Taken together, a panel of mutations within a hotspot region of ATP3 convey a strong fitness advantage and suppress the slow-growth phenotype for cells without mtDNA. Notably, we isolated exclusively ATP3 alleles as suppressors of petite growth during the adaptive evolution experiment that was run in replicates, with a strong selection pressure to optimize growth in minimal F1 medium, over hundreds of generations, and sequenced with high coverage. We hence conclude that ATP3 mutations are the most efficient solution to overcome petite growth, at least in the given genetic background and condition. Improved growth of evolved petites correlates with an increase in the mitochondrial membrane potential ΔΨ[mito] Strains suppressing the petite phenotype have an increased mitochondrial membrane potential (ΔΨ[mito]) compared to naive petites ^[133]29,[134]36,[135]48,[136]49 . Consequently, we examined changes in ΔΨ[mito] in strains carrying newly identified ATP3 alleles, and tested, to which extent these relate to the changes in growth rate. Using the MitoLoc reporter assay, we monitored differences in ΔΨ[mito], as well as mitochondrial morphology between naive and evolved petites, as well as in “reconstituted” evolved petites heterologously expressing the ATP3 alleles ^[137]50 . Wild type cells showed colocalization of the MitoLoc markers consistent with a functional ΔΨ[mito] ^[138]50 . Colocalization was lost in petites, indicating depolarisation ([139]Fig. 2a, left panels). Expression of the petite-suppressing ATP3 alleles restored mitochondrial colocalization of the MitoLoc markers, indicating an improvement of ΔΨ[mito] ([140]Fig. 2a, right panels). These results were corroborated by a quantitative colocalization analysis ^[141]50 ([142]Fig. 2b), and through staining with the ΔΨ[mito]-dependent dyeDiOC[6] ([143]Fig. 2c, [144]Suppl. Fig. 6). Similarly, ATP3-mutants isolated from the evolution experiments revealed an increase in ΔΨ[mito], compared to the petite controls ([145]Suppl. Fig. 7a). Overall, we found that the increase in ΔΨ[mito] ([146]Fig. 2b, [147]Suppl. Fig 7a) correlated with the gain in growth rate (R^2 = 0.93, [148]Suppl. Fig. 7b), suggesting a link to the restored growth phenotype of evolved petites ([149]Fig. 1e, f). Figure 2. Suppressor mutations increase the mitochondrial membrane potential, but do not restore mitochondrial morphology or metabolic capacity in petites. [150]Figure 2 [151]Open in a new tab A - C. The mitochondrial membrane potential is compromised in petites, but improves upon expression of ATP3-6 and ATP3-7. (A) Indicated YSBN11 strains were transformed with pMitoLoc^ [152]50 , a reporter construct containing two fluorescent proteins with ΔΨ^mito-dependent (preCOX4-mCherry) or ΔΨ^mito-independent (mtGFP) mitochondrial import. Naive petites accumulate preCOX4-mCherry in the cytosol and nucleus, indicative of reduced ΔΨ^mito. In contrast, in wild type (ρ^+) and petites (ρ^0) expressing ATP3 variants both markers colocalize, indicating ΔΨ^mito polarization. Scale bar = 1 μm. (B) Quantitative analysis of ΔΨ^mito estimated from Pearson Correlation Coefficients (PCC) of pixel-by-pixel protein colocalization as shown in (A). Loss of mtDNA leads to a significant decrease of ΔΨ^mito (61 %) compared to the wild type, which is significantly rescued by presence of plasmid-encoded ATP3 variants (up to 77 %). *p < 0.05, n > 46 cells, mean values +/- SEM. (C) Analysis of ΔΨ^mitoin the strains described in (A) as estimated by DiOC[6] staining followed by quantification with flow cytometry of n = 19956 (ρ^+), n = 16104 (ρ^0), n = 19112 (ρ^0 ATP3-6), n = 18893 (ρ^0 ATP3-7) events. D. Mitochondrial ATPase activity is reduced in strains without mtDNA. Mitochondrial proteins of YSBN 11 strains with indicated genotypes were isolated and ATPase activity was determined. Petites suffer from a significant reduction of the oligomycin-insensitive ATPase activity compared to wild type controls. Expression of variant ATP3 alleles did not rescue ATPase activity. Mean values of n = 5 biological replicates +/- SEM. Statistics is based on two-sided, unpaired t-tests comparing wild type against petite and evolved petite strains. *p < 0.05. E.-H. The metabolic capacity to metabolise glucose is unaffected in evolved petites. Indicated YSBN11 strains were cultured in batch fermenters for 45 hours in F1 medium containing 2 % glucose, and biomass as well as glucose, glycerol and ethanol concentrations were monitored. Wild type (ρ^+) cultures reach stationary phase after 13.9 h, petites (ρ^0) after 20 h, and petites expressing ATP3-6 and ATP3-7 variants after 15.9 h (dashed lines). (E) Biomass as estimated from OD[600] measurements over time. Wild type cultures accumulate 10.2 OD[600] units of biomass, while petite (6.1 OD[600] units) and evolved petites (6.1 and 6.4 OD[600] units) accumulate less biomass. After the diauxic shift, only wild type cultures continue to accumulate biomass. (F) Glucose levels in culture media. Note that glucose is entirely consumed during exponential growth. (G) Ethanol and levels in culture media. All strains produce similar levels of ethanol (up to a concentration of 7.7 - 8.0 g/L), while petites are unable to consume ethanol subsequent to glucose exhaustion. (H) Glycerol levels in culture media. Wild-type produces 0.78 g/L glycerol during exponential growth, which is subsequently consumed. Instead, petites produce higher levels of glycerol (1.7-1.9 g/L) during exponential phase, but are unable to consume glycerol after glucose exhaustion. While mitochondria in unstressed wild-type yeast cells are highly connected, petites exhibit a fragmented mitochondrial network, indicative of mitochondrial fission as it is induced by depolarisation of the mitochondrial membrane potential ^[153]50,[154]51 . Monitoring mitochondrial morphology using super resolution microscopy, we tested whether mitochondrial fission in petites is explained by their compromised ΔΨ[mito]. However, the mitochondrial network of petites carrying the ATP3 mutations displayed a similar degree of fragmentation compared to naive petite cells ([155]Fig. 2a, [156]Suppl. Fig. 8), indicating that depolarization of the mitochondrial membrane potential is not the only cause of mitochondrial fission in petites. Evolved petites grow faster than naive petites but form a similar amount of biomass A primary function of the respiratory chain, oxidative ATP biosynthesis, is interrupted in petites. Moreover, it has been hypothesized that in petites, the remaining and assembled F1 subunit of the ATP synthase ^[157]52 could be running in “reverse” to hydrolyse mitochondrially imported ATP, and in this way restore ΔΨ[mito] via the electrogenic exchange of mitochondrial ADP^3- with cytosolic ATP^ [158]4–[159]38,[160]41,[161]53–[162]57 . In both scenarios, a decreased ATP production could become growth limiting ^[163]31 . However, we found that the adenylate energy charge ^[164]58 was unaffected in petites and evolved petites compared to wild type cells ([165]Suppl. Fig. 9). ATP levels in petites only collapsed after glucose exhaustion ([166]Suppl. Fig. 10), when due to the absence of the respiratory chain, the petites can not consume the remaining non-fermentable carbon substrates, i.e. ethanol and glycerol ^[167]59 . Further, quantification of the Complex V ATPase activity revealed that it was strongly reduced rather than increased in petites and evolved petites compared to wild type ([168]Fig. 2d). Hence, we concluded that the evolution of fast growth in petites was not achieved through altering ATP metabolism, nor through a futile metabolic cycle in which increased ATP hydrolysis would restore the ΔΨ[mito.]. Next, we performed fermentation experiments to determine the efficiency of petite and evolved petite cells in converting the glucose into biomass. We determined carbon balances of wild type, petite, and evolved petites by measuring glucose uptake, ethanol and glycerol production, CO[2] and O[2] levels in the bioreactor, as well as the obtained biomass during growth on glucose ([169]Fig 2 e-h, [170]Suppl. Fig. 11). Further, we used the obtained data on glucose consumption, ethanol and glycerol production as well as obtained biomass in a model to estimate physiological parameters such as yield coefficients, growth rate, and uptake/excretion rates during the exponential phase of all strains (first 10.5 hours, [171]Suppl. Table 2). Wild type yeast consumed glucose and entered diauxic shift after 13.9 hours, having accumulated 168 % of the biomass generated by petites ([172]Fig. 2e). Petite cells reached glucose exhaustion after 20.0 hours ([173]Fig. 2f), reflecting their slower growth rate. Strainscarrying ATP3-6 or ATP3-7 reached stationary phase and glucose exhaustion considerably faster than petites (15.9 hours; [174]Fig. 2e, f). Despite accelerating growth, the ATP3 adaptationsdid not largely affect the biomass yield, i.e. the biomass produced per glucose consumed during exponential phase ([175]Suppl. Table 2), or the biomass obtained until the diauxic shift ([176]Fig. 2e), nor the ability of the petite cells to consume non-fermentable carbon sources ([177]Fig. 2g,h, [178]Suppl. Table 2). Petite suppression does hence not affect the general metabolic properties of petites. Instead, these results indicated that growth of petites is compromised because biosynthetic pathways do not meet the required rates to support faster growth. Insufficient supply of four amino acids limits growth of cells without mtDNA In order to pinpoint the biochemical processes affected in petite cells, we performed untargeted proteomics using microLC-SWATH-MS, a proteomic method designed for obtaining precise enzyme quantities in large sample series ^[179]60 . The proteomic data indicated that petites suffer from a broad range of metabolic perturbations that revolve around central carbon metabolism, the tricarboxylic acid (TCA) cycle in particular, and amino acid metabolism ([180]Fig. 3a and [181]Fig. 3b, squares). Although a contribution of the TCA cycle was expected due to the loss of crucial components of the respiratory chain, the degree of perturbed amino acid biosynthesis pathways was somewhat surprising. Curiously, recent findings from another yeast, Schizosaccharomyces pombe, which is normally petite negative, indicated that that supplementation with amino acids derived from α-ketoglutarate supported growth upon pharmacological inhibition of the respiratory chain ^[182]61 , supporting the hypothesis that in respiratory deficient cells, metabolism of α-ketoglutarate and downstream amino acids are challenged ^[183]62 . Figure 3. Amino acid metabolism is perturbed in petites, and their growth rate defect is rescued by the addition of selected amino acids. [184]Figure 3 [185]Open in a new tab A. Pathway enrichment analysis of proteomics data indicates metabolic defects associated with petites. Differentially expressed proteins between wild type (ρ^+) and naive petite (ρ^0) YSBN1 yeast were determined using SWATH-MS and subjected to pathway enrichment analysis using String-db. 11 Reactome pathways were significantly enriched (FDR < 5 %), with most significantly enriched pathways being biological functions related to TCA and amino acid synthesis. Enrichment is shown as bar size (-log10 enrichment p-value) and bar colors indicating enrichment over background frequency (number of changed proteins compared to all proteins of that pathway), and 9 TCA pathway proteins (enrichment = 0.56) were changed in naive petites. Protein abundance data from n = 4 biological replicates per genotype. Note that the p-values stated in the corresponding barchart represent non-log transformed values. B. Differential expression of proteins and metabolites of naive and evolved petites compared to wild type. A simplified pathway representation of central carbon metabolism in S. cerevisiae. Mean fold-changes of enzyme (squares) and metabolite levels (circles) of petite (middle column, empty plasmid control) and evolved petites (right column, expressing ATP3-6) against wild type (left column, empty plasmid control) are shown. Data shown is mean fold change from n = 4 biological replicates per genotype. C-E. The slow growth of petite cells is rescued by a supplementation with glutamate, glutamine, arginine and leucine. (C) Growth rate of YSBN11 wild type (containing empty vector control), naive petites (containing empty vector control) and evolved petites (containing ATP3-6) grown in minimal media with 2 % glucose in 96-well liquid culture, without (-AA) or with addition with a complement mixture of proteinogenic amino acids (+AA), or combinations of 2 mM glutamate (E), 2 mM glutamine (Q), 2 mM arginine (R) and 2 mM leucine (L). The slow-growth phenotype (0.075 OD min^-1) of petites increases to 0.137 OD min^-1 upon supplementation with QERL. Mean values +/- SD of n = 5 biological replicates. (D) The relative growth defect of petite cells (normalized to percentage of wild type growth rate) is compensated from 43.3 % to 13.5 % upon QERL supplementation, as calculated from the data as shown in (A). Mean values +/- SD of n = 5 biological replicates. (E) Indicated strains were pre-grown in minimal media with (+QERL) or without supplements (-AA). A serial dilution series (1:5) was spotted with an initial OD[600] of 0.5 on agar plates with the indicated media composition. Shown is one of n = 2 biological replicates. We hence continued with a quantification of amino acids and respective precursors, which confirmed perturbation of intermediary metabolism in petites ([186]Fig. 3b. circles). Petites were characterized by strongly reduced concentrations of leucine, arginine (as well as its precursors citrulline, ornithine and α-ketoglutarate). In parallel, enzymes of the associated synthesis pathways were differentially expressed in petites. Upon expression of the mutant ATP3 alleles, these expression changes were partially alleviated ([187]Fig. 3b squares, [188]Suppl. Fig. 12). Curiously, the amino acid levels did not follow the expression changes, and remained low in evolved petites compared to wild-type cells ([189]Fig. 3b, circles). We speculated that this result could indicate that the demand for these metabolites was indeed limiting. We therefore conducted supplementation experiments with combinations of leucine, arginine, as well as glutamate and glutamine, which both are directly associated with α-ketoglutarate metabolism. Supplementation of these amino acids, in separation, and specifically in their combination, lead to a disproportionate increase in the growth rate of petites compared to how wild type or evolved petites responded to supplementation ([190]Fig. 3c, [191]Suppl. Fig. 13). Consistent with previous observations, leucine shortages are particularly growth limiting due to the high biosynthetic requirements for this amino acid ^[192]63–[193]65 , compensating with leucine had a particularly strong impact on the petite growth rate ([194]Fig. 3c). The supplementation of arginine, which has also been described as beneficial to petites in another study ^[195]66 , but also the combination of glutamine and glutamate benefited petite growth clearly more than it benefited wild type or evolved petites. When we supplemented all four amino acids (QERL) at once, the growth rate of petite cells became similar to the growth rate of wild-type cells without amino acids (AA) supplementation ([196]Fig. 3c, d, e). Supplementation with all proteinogenic AA also accelerated growth, but the complex treatment was less effective than the addition of just QERL alone ([197]Fig. 3c). The lower efficacy of the broad rather than the specific supplementation is potentially caused by multiple amino acids competing for the same transporters ^[198]67 , or by substrate-induced down-regulation of amino acid transport in more complex media ^[199]68–[200]71 . Further, we tested if improved growth by QERL supplementation coincides with an increase in ΔΨ[mito], that is known to depend on growth rate. Indeed, the membrane potential was generally higher in the fast growing cells. However, differences between wild type, petite, and evolved petite overall remained ([201]Suppl. Fig. 14). Taken together, these data suggest that insufficient rates in the synthesis of specific amino acids, especially of leucine, arginine, glutamine and glutamate, is limiting the growth rate of petites. Perturbations of NADPH metabolism, TOR signalling, and the retrograde response are not causal for the impaired amino acid biosynthesis in petites The observation that multiple amino acid biosynthesis pathways were affected in parallel indicated that regulatory processes, the cellular-chemical environment, or common metabolic precursors, or combinations of these, were limiting the growth in petite cells. First, we tested if an amino acid deficiency is associated with a lack of reducing equivalents in the form of NADPH, which is required for amino acid anabolism ^[202]72,[203]73 . We exploited the phenotype that yeast cells are sensitive to the synthetic thiol oxidant diamide when NADPH supply is limited, as NADPH is also providing reducing power for the antioxidant machinery ^[204]73–[205]76 . We did not detect strong differences in petites in their tolerance to diamide ([206]Suppl. Fig. 15a). Further, we conducted measurements of the glutathione (GSH) pools, that depend on NADPH availability. GSH pools were comparable between wild type and petite strains ([207]Suppl. Fig. 15b–[208]d). Finally, we measured superoxide levels. These were marginally changed, and indeed, we detected lower superoxide levels in petites than in wild-type cells ([209]Suppl. Fig. 15e). Taken together, these results argue against a significant NADPH shortage or strong oxidative imbalances as a cause of the amino acid deficiency of petite cells. Next, we focussed on amino acid sensing and signalling. We studied the response of petites and evolved petites to TOR inhibition by rapamycin ^[210]77 . A dose response curve revealed that the chemical sensitivity (IC50) to rapamycin does not differ between cells with or without mtDNA ([211]Suppl. Fig 16). We then investigated the role of mitochondrial retrograde (RTG) signalling, which has been attributed to mitochondrial dysfunction ^[212]1 . In addition to regulating the expression of mitochondrial genes such as the TCA cycle, the RTG response has been implicated in regulation of peroxisomal processes ^[213]2 . As peroxisomal proliferation has hence been associated with mitochondrial deficiency ^[214]78 , we used a peroxisomal fluorescent marker (Pts1p-GFP) to quantify these organelles. The number of peroxisomes was significantly increased in petites compared to the wild type (18.5 ± 1.0 vs. 13.0 ± 0.6 in WT; p < 0.0001). Moreover, the number of peroxisomes returned close to wild type levels in evolved petites (11.2 ± 0.6; p = 0.05) ([215]Suppl. Fig. 17a–[216]b). Similarly, differential expression of proteins associated with the RTG regulon, Aco1p, Dld3p, Idh1p, Idh2p and Pyc1p ^[217]1,[218]79 was observed in petites, and this response was alleviated in evolved petites([219]Suppl. Fig. 17c).These results indicated that the RTG response pathway is activated in petites, but not in evolved petites. To test whether suppression of retrograde signalling was causally associated with the amino acid deficiency, we deleted RTG2, the central mediator of the retrograde response, and CIT2, a target of a downstream regulated process, the glyoxylate cycle ^[220]1 . Deletion of RTG2 or CIT2 did not restore petite growth, instead, the growth defect in petites increased. Moreover, rtg2 and cit2 mutants did not affect the ability of ATP3 mutants to suppress the petite phenotype ([221]Suppl. Fig. 17d). These results suggest that activation of the retrograde response pathway is a protective response for cells without mtDNA, but not causing the petite phenotype. Two enzymatic steps of the TCA cycle are inhibited in petite cells The retrograde signalling is typically associated with the synthesis of α-ketoglutarate and downstream amino acids ^[222]62 . We therefore focused on the TCA cycle, of which α-ketoglutarate is a key intermediate. Quantification of both proteins and metabolites involved in TCA cycle reactions revealed that the upper TCA cycle is severely affected in petites compared to wild type cells ([223]Fig. 4a, b, [224]Suppl. Fig. 18, 19). Pathway intermediates including citrate, aconitate and succinate accumulated, whereas the concentration of α-ketoglutarate was reduced ([225]Fig. 4b). Figure 4. Aconitase is inhibited inpetites and blocks the TCA cycle, which is alleviated in evolved petites. [226]Figure 4 [227]Open in a new tab A-B. Differentially expressed proteins and metabolites of the TCA cycle. A simplified pathway representation of central carbon metabolism in S. cerevisiae. Shaded areas indicate respective cellular localization (mitochondrial, peroxisomal, cytosolic) labelled according to the small inlet (top right). (A) Enzymes (squares) are indicated next to associated reactions (arrows). Mean fold-changes of protein levels (squares) of wild type YSBN11 (left column, plasmid control) compared to petites (middle column, plasmid control) and an evolved petite strain (right column, expressing plasmid based ATP3-6) are shown. Note that electron transport chain and succinate dehydrogenase are impaired in petites ^[228]1,[229]7,[230]8 . Associated enzymes are accordingly dysregulated, and not alleviated in evolved petites. Instead, the reactions from citrate to α-ketoglutarate are differentially regulated in petites, but partially restored in evolved petites. Shown are mean values of n = 4 biological replicates per genotype. (B) Mean fold-changes of metabolite steady-state levels (circles) indicated next to the associated reaction in the pathway. Genotypes and data normalization like in (A). Metabolite levels indicate two blocks in the TCA cycle at the steps of aconitase and succinate dehydrogenase. Shown are mean values of n = 3 biological replicates per genotype. C. Aconitase and isocitrate dehydrogenase protein levels are increased in petites. Protein expression illustrated relative to the wild type (ρ^+) strain highlight upregulation of aconitase and isocitrate-dehydrogenase in petites. Levels of the major NAD- and NADP-dependent isocitrate dehydrogenase (Idh2p/Idp1p respectively), as well as the main mitochondrial aconitase (Aco1p) are upregulated in petites. Impairments are partially alleviated in an evolved petite strain. Mean values +/- SD of n = 3 biological replicates. Statistics based on two-sided, unpaired t-tests comparing wild type to other genotypes. Note that the data is from the same experiment as shown in (A). D. Aconitase is inhibited, while (NAD dependent) isocitrate-dehydrogenase activity is increased in petites. Activity of the indicated enzymes was quantified using enzyme assays with indicated yeast extracts and normalized to the wild type (ρ^+) strain. Despite the increased expression level (C), Aconitase activity was reduced below the detection limit in petites, and regained activity in evolved petites. In contrast, NAD-dependent isocitrate dehydrogenase activity mirrors the determined protein levels (shown in C) and increases in activity in petites. Mean values +/- SD of n = 3 biological replicates per genotype. Statistics based on two-sided, unpaired t-tests. E. Major carbon fluxes in wild-type, petites, and evolved petites during exponential growth on glucose. Wild-type (top row), petite (middle row), and evolved petite (bottom row) were grown in bioreactors (corresponding to [231]Fig. 1) and fermentation products were monitored. Shown are the estimated carbon balanced - glucose uptake rate, ethanol and glycerol production rate, as well as calculated remaining carbon flux allocated to CO2 that is indicated as the TCA cycle during exponential growth phases. Top row corresponds to wild-type, middle row to petite, and bottom row to evolved petite. Indicated rates are given as c-mol / (gDW*h) with six carbon equivalents for glucose, three for glycerol, and three for ethanol (assuming stoichiometric production of one CO2 per mole of ethanol). Shown uncertainties were calculated from the model based estimations. For corresponding data, see [232]Suppl Table 2. F-G. Stable isotope labelling incorporation into and levels of pyruvate, TCA cycle, and glutamate, glutamine, GABA. YSBN11 wild type, petite, and evolved petites carrying plasmid-encoded ATP3-6 as indicated were grown exponentially, harvested, and resuspended in media with [U-^13C]-glucose. After 1, 5, 15 and 30 min cells were quenched, extracted, and analyzed using GC-MS. (F) Fractional labelling (estimated as percent of the metabolite pool containing ≥ one ^13C carbon atoms) is similar between genotypes. (G) Pool sizes (estimated from the sum of all isotopologues) of indicated metabolites differ between genotypes. While wild type and evolved petites display largely similar levels of indicated metabolites, petites accumulate citrate/cis-aconitate, whereas pool sizes of glutamate, glutamine and GABA are reduced. Shown are mean values +/- SD of n = 3 (wild type, evolved petite), n = 2 (petite) biological replicates. A partial explanation for the perturbation of the TCA cycle in petites is its connection with the respiratory chain. The loss of the mitochondrial genome deletes Cox1p, Cox2pandCox3p that are implicated in electron transport from succinate dehydrogenase (SDH, complex II) to cytochrome c reductase (complex III) ^[233]1,[234]7,[235]8 . Without a functional succinate dehydrogenase, and potentially fostered by activation of the glyoxylate shunt, succinate accumulates in petites ([236]Fig. 4b). While our data is consistent with a loss of SDH activity, it did reveal a second and so far overlooked perturbation within the TCA cycle ([237]Fig. 4b). Levels of citrate and cis-aconitate were increased in petites, while the α-ketoglutarate pool was reduced ([238]Fig. 3b). On the protein level, α-ketoglutarate dehydrogenase (Kgd1p, Kgd2p) and succinyl-CoA ligase (Lsc1p, Lsc2p) were down-regulated, while aconitase (Aco1p, Aco2p) and especially isocitrate dehydrogenase (Idh1p, Idh2p, Idp1p, Idp3p) were up-regulated in petites ([239]Fig. 4a, c). Upregulation of aconitase and isocitrate-dehydrogenase isomers in petites contrasted to the accumulation of their respective substrates (citrate and cis-aconitate). Moreover, enzymes involved in catalysis of citrate to α-ketoglutarate returned to wild type levels in evolved petites. These results implied that one or more of the upper TCA cycle enzymes could be less active specifically in petites, and therefore be upregulated as a compensatory response.We quantified enzyme activities from wild-type, petite, and evolved petite extracts using enzyme assays to monitor substrate conversion ([240]Fig. 4d). NAD-dependent isocitrate dehydrogenase (IDH) activity was increased in petites according to the increased expression levels of the enzyme (Idh1p and Idh2p, [241]Fig. 4a, c, d). The NADP dependent isocitrate dehydrogenase was unchanged ([242]Fig. 4d). In sharp contrast, despite the increase in aconitase protein expression in petites ([243]Fig. 4c), the activity of this enzyme was reduced below the detection limits of the assay ([244]Fig. 4d). Remarkably, aconitase activity was again detected in the evolved petites ([245]Fig. 4d). We asked whether the inhibition of the TCA cycle enzymes is reflected in lower activity of this cycle. Subtracting the carbon balanced production rate of the major fermentation products ethanol and glycerol from the glucose uptake rate allowed us to estimate the major carbon fluxes between wild-type, petite, and evolved petites. Indeed, considering their biomass yield and growth rate, petites channel more carbon from glucose to ethanol/glycerol compared to wild-type ([246]Suppl. Table 2, [247]Fig. 4 e). This also suggests less carbon is allocated to the CO[2] producing TCA cycle. Then, we performed ^13C tracing experiments to estimate the relative activity of the TCA cycle. We followed labelling incorporation of [U-^13C]-glucose to pyruvate, and through the upper TCA cycle to succinate ([248]Fig. 4 f, g, [249]Suppl. Fig. 20). Of note, as both metabolomics and a metabolite labelling based strategy can not distinguish shared pathway intermediates of the mitochondrial TCA cycle with the cytosolic/peroxisomal glyoxylate cycle, these data have to be seen as a mix of both signals. Moreover, as a surrogate for α-ketoglutarate, which could not be quantified in petites during tracer experiments, as by addition of labelling, the signals were too diluted, we analysed label incorporation into and pool sizes of glutamate, glutamine, and GABA as downstream metabolites. While labelling incorporation after 30 min was largely similar, the substrate levels of aconitase, citrate and cis-aconitate were elevated. Instead, the downstream metabolites glutamate/glutamine/GABA were reduced in petites, but not in evolved petites ([250]Fig. 4f, g). Especially considering that petites have less carbon flux ([251]Fig. 4e), these data are consistent with a reduced activity of reactions towards glutamate/glutamine/GABA specifically in petites. In contrast, succinate levels were elevated in petites but also in evolved petites ([252]Fig. 4b, Fig. 4g), which is in line with the loss of SDH activity in all strains lacking mtDNA. In summary, aconitase, isocitrate dehydrogenases and the glyoxylate shunt, targets of the retrograde response, are upregulated in petites. Petites are further characterized by strong reduction in aconitase and SDH activity, while evolved petites overcome the inhibition of aconitase. Loss of mtDNA increases intracellular iron levels and is associated to aconitase inhibition in petites Next, we addressed potential mechanisms that could inhibit aconitase activity in petites. Aconitase is a prominent iron-sulfur-cluster (ISC) dependent enzyme ^[253]80 . Defects of mtDNA and ΔΨ[mito] have been implicated with perturbed iron- and ISC-homeostasis ^[254]30,[255]81–[256]83 . We hence speculated that an impaired iron /and ISC metabolism could be responsible for the loss of aconitase activity. First, we tested for the activation of the iron regulome ^[257]84,[258]85 by recording proteomes of cells depleted and replete with iron ([259]Fig. 5a, b). In order to distinguish confounding effects caused by the differences in growth rate we applied principal component analysis (PCA, [260]Suppl. Fig. 21). Growth-rate (PC1) and iron depletion (PC2) response were the two dominant drivers of proteome changes. Growth-rate sensitive proteins (PC1) changed substantially in petites, but not in wild type and evolved petites ([261]Fig. 5a). Instead, the top 50 iron-response proteins (PC2) changed according to iron depletion in all wild type, petite, and evolved petite cells ([262]Fig. 5b). This includes proteins expected to change in response to iron depletion, such as ISC proteins that are down-regulated upon iron starvation^ [263]86 ([264]Fig. 5b red dots, [265]Suppl. Fig. 22). Thus, petites mount a functional response to iron depletion. Figure 5. Iron overload inhibits aconitase, which is required for petite suppression. [266]Figure 5 [267]Open in a new tab A-B. The iron starvation-response is activated in petites. YSBN11 wild type (plasmid control), petite (plasmid control), and an evolved petite (expressing plasmid-based ATP3-6) were cultured with (+) or without (-) iron supplementation as indicated. Protein abundance was determined by SWATH-MS, and fold-changes were calculated in comparison to iron-replete (+iron) wild type and analyzed by principal component (PC) analysis ([268]Suppl. Fig. 21). (A) Heatmap displaying the 50 proteins with highest loadings in PC1 (associated with growth rate). Irrespective of iron addition, levels are consistently up regulated in petites, compared to both wild type or evolved petitesexpressing ATP3-6. (B.) Iron response mounts irrespective of the genotype. Heatmap illustrating the 50 proteins with highest loadings in PC2 (associated with iron-depletion). ISC proteins (red dots) are severely down regulated; levels of an iron transporter (blue dot) increases. Data from n = 4 biological replicates. C. Elevated cellular iron levels in petites. Iron levels in a YSBN11 wild type (plasmid control), a petite (plasmid control), and an evolved petite strain (expressing plasmid-based ATP3-6) were analyzed by ICP-MS and normalized to protein content. Cellular iron is increased in petites (0.17 nmol Fe/μg protein) compared to wild type (0.81 nmol Fe/μg) and evolved petites (0.80 nmol Fe/μg). Mean values +/- SD of n = 3 biological replicates. Statistics based on unpaired, two-sided t-tests. *p < 0.05. D. Petites are more resistant to iron depletion compared to wild type. (D) Dose response curve of YSBN11 wild type (black), petite (red), and evolved petites expressing ATP3-6 (blue). Shown is the normalized growth rate against increasing concentration of the iron chelator DIP. Growth rate was normalized to the maximum growth rate obtained with physiological iron concentration of the respective genotypes, as indicated by the dashed line. Mean values are indicated by lines, and standard deviation by area. Data from n = 6 biological replicates. E. Aco1 binds excessive iron in petites. BY4741 wild type (expressing Aco1-GFP, black), a petite (expressing Aco1-GFP, red), and an evolved petite (expressing Aco1-GFP and ATP3-6, blue) were exposed to ^55Fe. Aco1-GFP was enriched from cell extracts, and bound ^55Fe was quantified by scintillation counting. While GFP alone bound neglectable amounts of iron (18 cpm ^55Fe/mg protein), Aco1-GFP incorporated 636 cpm (wild type), 6734 cpm (petite), and 1296 cpm (evolved petite) of ^55Fe/mg protein respectively. Thus, Aco1p binds excessive iron specifically in petites. Statistics based on unpaired, two-sided t-tests of n = 3 biological replicates. *p < 0.05. F. Mitochondrial aconitase activity is reduced in petites. Wild type and petite BY4741 yeast were transformed with plasmid controls, or plasmids harbouring native aconitase (ACO1) or additionally fused to a mitochondrial signaling sequence (mtACO1) respectively. In empty vector controls, aconitase activity was detected in wild type (6.9 mU/mg), but not in petites (0.5 mU/mg). Heterologous expression of ACO1 and mtACO1 increased aconitase activity in wild type (16.7 mU/mg and 13.1 mU/mg). In petites, only ACO1 expression increased aconitase activity (5.4 mU/mg), but not when targeted as mtACO1 (0.6 mU/mg). *p < 0.05, n=3. G. Aconitase is sensitive to increasing iron levels in vitro . Extracts of YSBN11 wild type (ρ^+, plasmid control) were incubated with increasing concentrations (0 - 500 μM) of Fe[II]-chloride and aconitase activity was determined. Exposure to iron concentrations ≥ 50 μM significantly reduced aconitase activity. Mean values are indicated by lines, and standard deviation by the grey area. Statistics based on unpaired, two-sided t-tests of n = 3 biological replicates per time-point. *p < 0.05. H. Iron-dependent inhibition aconitase is alleviated by addition of the iron chelator deferoxamine. Genotypes and experimental setup as described in (E). Cell lysates were incubated with or without 100 μM iron[II] chloride and 20 mM deferoxamine (DFO-B) and assayed for aconitase activity. While addition of 100 μM iron led to a significant reduction in aconitase activity, addition of 20 mM DFO-B alleviated iron-induced enzyme inhibition. Statistics based on unpaired, two-sided t-tests of n = 3 biological replicates per time-point. *p < 0.05. I. Petites no longer benefit from expression of ATP3-6 upon deletion of aconitase. Growth of indicated strains (genotypes and color code like in (G)) was monitored and growth rates were determined. The slow-growth phenotype of petites was rescued by expression of ATP3-6, but not in a Δaco1 background. Statistics based on unpaired, two-sided t-tests from n = 3 biological replicates. *p < 0.05. J. Proteome profile of Δaco1 does not respond to expression of ATP3-6 . Levels of 1929 proteins were determined in BY4741 wild type (ρ^+) expressing plasmid control (black) or ATP3-6 (grey), petites (ρ^0) expressing plasmid control (red) or ATP3-6 (blue), and petite Δaco1 expressing plasmid control (yellow) or ATP3-6 (orange). A PCA was performed with normalized intensities of all proteins. Proteome profiles of wild type (black, grey) cluster closely together, while slow- (red) and fast-growing (blue) ρ^0 strains are well separated along PC2. In contrast, ρ^0 Δaco1 (yellow) is not separated from Δaco1 expressing ATP3-6 (orange). Data from n = 3 biological replicates. Next, we tested whether excessive ion concentrations could be a problem. We conducted a total elemental composition analysis to quantify a range of elements including iron in wild type, as well as naive and evolved petites by inductively coupled plasma-mass spectrometry (ICP-MS, [269]Suppl. Table 3). We found iron levels were significantly elevated (> 2-fold) in petites compared to the wild type. Other element concentrations were not affected. ([270]Fig. 5c, [271]Suppl. Table 3). Strikingly, iron levels returned to wild type upon expression of the petite suppressor ATP3-6 ([272]Fig. 5c). In order to test whether the highly elevated iron levels could be associated with petite growth, we first manipulated growth media using iron chelators. The relative growth rate of wild type cells dropped with increasing iron chelator concentrations. Relative to their slower growth rate, petites did better tolerate iron depletion ([273]Fig. 5d, [274]Suppl. Fig. 23). To test if aconitase might be directly affected by elevated iron levels in petites, we supplemented the cells with the radioactive iron isotope ^55Fe, and affinity-purified mitochondrial Aco1p-GFP. Protein-bound iron from immuno-purified Aco1p-GFP wasquantified by scintillation counting. Aconitase enriched from wild type yeast showed weak radioactivity. Instead, aconitase purified from petites bound a high degree of radioactivity ([275]Fig. 5e). We then heterologously overexpressed ACO1, which is localized in both cytoplasm and mitochondria^ [276]87 . Furthermore, we overexpressed ACO1 fused with an additional mitochondrial targeting signal ^[277]50 . In a wild type background, aconitase activity was detected, and overexpression of both ACO1 constructs increased aconitase activity consistent with its overexpression ([278]Fig. 5f). Instead, in petite extracts, aconitase activity was reduced to the detection limit of the assay. Moreover, in the petite, no increase in total aconitase activity was detected when the overexpressed protein was specifically targeted to mitochondria ([279]Fig. 5f). Aconitase activity could be low because it is not correctly folded/assembled (e.g. not correctly loaded with it`s ISC). Furthermore, a principally functional enzyme could be sensitive to high iron levels. To test this, we used wild-type extracts, and tested aconitase activity in vitro, with an increasing iron concentration. As expected, in wild-type extracts, aconitase was active. However, the activity decreased with increasing iron concentrations ([280]Fig. 5g). Adding the iron chelating agent deferoxamine (DFO-B) prevented this effect ([281]Fig. 5h). Nonetheless, wild-type aconitase retained a higher activity in the presence of iron, compared to aconitase purified from petites ([282]Fig. 5g, f). These results imply that two scenarios apply: aconitase purified from petites is dysfunctional and has bound excessive iron ([283]Fig. 4 d, [284]Fig 5e, 5f), but also, a per se functional aconitate is sensitive to increased iron levels ([285]Fig 5g, h). These findings suggested that aconitase itself would be required to suppress the petite phenotype. To test this, we generated an aco1 deletion strain. Δaco1 loses respiratory function and mtDNA ([286]Suppl. Fig. 24a) ^[287]88 , and is auxotrophic for glutamate ([288]Suppl. Fig. 24b), which is in line with perturbation of the TCA cycle reactions towards α-ketoglutarate ^[289]89 . Likewise, growth of petites further deteriorates upon deletion of aco1 ([290]Fig. 5i). Based on classification with PCA, petites with intact ACO1 have a strong shift along PC2 upon expression of ATP3-6 ([291]Fig. 5j, [292]Suppl. Fig. 25). This effect was abolished upon aco1 deletion from petites, indicating that the suppressor mutation ATP3-6 lost its function when aconitase was disabled ([293]Fig. 5j). Moreover, petiteΔaco1 cells expressing mutant ATP3 could no longer suppress the slow-growth phenotype ([294]Fig. 5i). Thus, a functional aconitase is required for petite-suppression, and the inhibition of the Krebs cycle is directly associated with the slow-growth phenotype of petites. Discussion The loss of mitochondrial DNA in yeast causes a severe growth defect described in 1949 as the petite phenotype, whose molecular basis was never fully clarified ^[295]17 . Mitochondria are a major organelle implicated in many cellular processes ^[296]1–[297]6 , and consequently, the loss of the mitochondrial genome has broad consequences on yeast physiology. In order to pinpoint the processes whose interruption is causal for the slow growth phenotype, we harnessed the power of adaptive evolution. Remarkably, accelerated growth of petites was readily evolved, and was explained by a set of mutations located in the C-terminal region of Atp3p, the ɣ subunit of the Complex V. Mutations in this hotspot region are known as one of the most common substitutions in the yeast genome ^[298]32,[299]36,[300]42,[301]45,[302]46 . We recently provided evidence that the frequent occurrence of such mutations are coinciding with an increase in the spontaneous mutation frequency, once the mitochondrial genome is deficient ^[303]32 . Our adaptive evolution experiment adds to these studies, by explaining that the selective advantage of ATP3 alleles is provided by maximizing growth rate, and by demonstrating that ATP3 mutations are a main solution to restore growth in petite yeast. This is because the design and long-term nature of our experiment would have allowed a number of possible adaptations to be selected, but a series of and independently emerging alleles all affected the ATP3 locus. In petites, the ATP3 mutations are dominant and similar to previously isolated mutant alleles that restore the mitochondrial membrane potential. Our data is effectively ruling out futile metabolic cycles to underlie the restoration of the ΔΨ[mito] because biomass formation does not change with petite suppression. It is thus possible that ATP3 suppressor mutations are gain of function mutations that generate a so far unknown biochemical activity. Although additional work will be required to elucidate the the mechanistic details, it is sensible to assume that elevated ΔΨ[mito] benefits mitochondrial functionality and thus various processes involving mitochondria, such as mitochondrial protein import and homeostasis ^[304]10,[305]90 . In this study, we focused on the physiology of these evolved petites, and used them as a model for identifying growth limiting factors of cells that have lost the mtDNA. We employed the basic assumption that perturbing the mechanism “repaired” in evolved mutants would leave the otherwise dominant allele dysfunctional. The effectiveness of this strategy is illustrated by ruling out retrograde response as a cause of the slow growth phenotype. We find that the RTG response is activated upon loss of the mitochondrial genome, and mitigated in evolved petites. However, the deletion of crucial components of the retrograde response impaired rather than restored petite growth. Further, the selected ATP3 alleles could still improve growth of petites in which the RTG was disrupted. Activation of the RTG is hence beneficial to petites as a downstream protective response, but cannot explain the growth benefit of isolated suppressor mutations. Eventually, our data converged in a model in which petites grow slowly because they are unable to synthesize amino acids, specifically glutamate, glutamine, arginine and leucine at a sufficient rate. We then continued to identify the key underlying molecular mechanisms. We discover that the TCA cycle in petites, which supplies biosynthetic precursors, or shares metabolites or iron sulfur cluster enzymes with the four metabolic pathways, is interrupted at two enzymatic steps, succinate dehydrogenase and aconitase. SDH is interrupted due to the loss of the mitochondrial genome ^[306]1,[307]7,[308]8 and remains deficient in the evolved petites. The inhibition of aconitase is overcome by evolved petites. Furthermore, once aconitase is deleted, the ATP3 mutations lose their function, and do no longer benefit petites. Our data indicates that two mechanisms contribute to the sensitivity of aconitase upon loss of mtDNA. First, we found that aconitase binds excessive amounts of iron in petites, and increased iron levels inhibit aconitase, even if it is functionally loaded with an ISC. Second, our results together with previous findings argue that in petites ISC assembly and export is perturbed due to a reduction of ΔΨ[mito] ^[309]30,[310]81–[311]83 . Reduced export of ISC can in turn explain elevated mitochondrial iron levels, and can contribute to incorrect loading of ISCs ^[312]81 . Of note, these results seem relevant for the pathology of Friedreich’s ataxia. This rare metabolic disease is associated with a mutation in the mitochondrial iron chaperone Frataxin, leading to accumulation of iron in mitochondria and aconitase deficiency ^[313]3,[314]91 . All four amino acids (QERL) that we uncovered to rescue the growth defect in petites rely on at least partial mitochondrial biosynthesis ^[315]92,[316]93 . Glutamate, which is in turn required for synthesis of glutamine and arginine, is directly affected by the perturbation of aconitase as its biosynthesis depends on α-ketoglutarate. Leucine ties into this perturbation as its synthesis requires Acetyl-CoA, and is inhibited by CoA ^[317]92,[318]94 . Another aforementioned common denominator in the biosynthesis of these amino acids is the requirement for intact iron sulfur clusters, specifically in the case of leucine (ILV3, LEU1) and glutamate/glutamine (GLT1) ^[319]81,[320]82,[321]92 . Nevertheless, as petites are viable, there must be residual ISC assembly, and in turn loading of ISC-dependent enzymes. Of note, yeast is particularly sensitive to perturbations in leucine metabolism. Leucine is one of the most frequently encoded amino acids, needs to be synthesized at high rates, and the strong growth defect of leucine auxotrophs cannot fully be compensated by supplementation ^[322]63–[323]65 . It hence seems plausible that a shortage in ISC metabolism would manifest in the leucine pathway, not only because it is in high demand, intertwined with the Krebs cycle and mitochondrial metabolism, but additionally because it depends on mitochondrial ISC export ^[324]95 . Taken together, while numerous cellular processes depend on mitochondria, the main growth rate restriction caused by mitochondrial genome loss is explained by a well-defined, biosynthetic defect in amino acid metabolism-most of it that can be overcome with a simple nutritional supplementation. These findings suggest that even complex mitochondrial dysfunctions, as they are seen in numerous genetic and acquired human diseases, could be overcome by mapping and compensating for their precise biochemical defects. Material & Methods Chemicals and media All chemicals were obtained from Sigma-Aldrich unless stated otherwise. Prototrophic yeast strains were cultured in F1 ^[325]44 or SM (6.8 g/L yeast nitrogen base (Sigma) medium, and growth of auxotrophic strains in SM was enabled by supplementing with 20 mg/L histidine, 60 mg/L leucine, 20 mg/L uracil, 20 mg/L methionine and/or 50 mg/L lysine as indicated. Synthetic complete (SC) medium was prepared by supplementing SM medium with CSM-HIS-LEU-MET-TRP-URA (MP Bio), 10 mg/L adenine, 20 mg/L uracil, 40 mg/L tryptophan, and amino acids as above. YP (20 g/L peptone (Bacto), 10 g/L yeast extract (Bacto) medium was used where indicated. Iron-free culture of yeast was performed using SM medium prepared with yeast nitrogen base without amino acids and without iron (Formedium), and analytical grade EMSURE water (Merck). Where indicated, iron-free medium was supplemented with 200 μg/L FeCl[3] 6 H2O. Media were supplemented with 2 % glucose unless stated otherwise. Δaco1 yeast strain was grown by additional supplementation with 5 mM glutamate. Yeast strains Adaptive evolution experiment was carried out using wild type YSBN1, a prototrophic diploid variant of S. cerevisiae S288c ^[326]44 . For transformation of yeast with plasmid constructs, derivatives of YSBN1 with single deletions of ura3 or his3 were used ^[327]44 . Where indicated, BY4741 strain was used. Single gene deletion mutants BY4741 Δcit2, Δzwf1 and Δrtg2 were obtained from the yeast deletion collection ^[328]96 and identity verified by sequencing. BY4741 Δaco1 was generated by inserting a natMX targeting cassette ^[329]97 into the ACO1 locus and verification by sequencing. Strains depleted of mtDNA (ρ^0) were generated by plating yeast on YPD agar containing 0.1 g/L ethidium bromide and incubation at 30°C for 2 d. Absence of mtDNA was confirmed in isolated surviving clones by DAPI staining and growth assays using the non-fermentable carbon sources ethanol and glycerol. For ρ^0 strains carrying a plasmid, yeast was transformed according to standard procedures, grown on selective agar and positive clones isolated. Single clones were then subjected to mtDNA depletion procedure as described above. Plasmid constructs Plasmids used in this study are listed in [330]Suppl. Table 4 and are available via Addgene. For construction of various pRS313_ATP3 plasmids, respective ATP3 variants (T911A, G919C, T911A+G919C) were first obtained by PCR site directed mutagenesis. Resulting alleles, and the respective native promoter region (the ~600 bp upstream region) were sub-cloned using the pGEM-T System (Promega). Subsequently, respective ATP3 alleles (including wild type) were cloned into the multiple cloning site of pRS313 ^[331]98 by digests with SalI/BamHI. Successful insertion was evaluated using blue/white selection. Finally, Sanger sequencing confirmed the respective ATP3 alleles of pRS313-ATP3 constructs and excluded additional mutations. Plasmid expressing Aco1p-GFP has been described elsewhere ^[332]99 , and plasmid expressing GFP-Pts1 was a gift from E. Hettema ^[333]100 . Construct for expression mitochondrially targeted aconitase was generated by sub-cloning of ACO1 and addition of preSu9 sequence using homologous recombination, and insertion into p416GPD using BamHI/SalI sites. Adaptive evolution and genome sequencing Wild type and petites were cultured in F1 carbon-limited medium ^[334]101 with 2.5 g/L glucose as carbon source in a DASGIP Parallel Bioreactor Systems chemostat for 67 days, with pH controlled at 4.5, agitation at 200 rpm, and temperature at 30°C. Three chemostat cultures were grown for each cell line a dilution rate of 0.1. Aliquots of each culture were archived daily with 25 % glycerol at -80°C. For genome sequencing, YPD was inoculated with glycerol stocks of chemostat samples and harvested in exponential phase. A 5 mL yeast culture was grown overnight in YPD and 1.4 mL culture was added to a microcentrifuge tube. Genomic DNA was extracted with a Wizard Genomic DNA Purification Kit (Promega) according to the manufacturer’s protocol. Whole-genome sequencing was performed on an Illumina HiSeq 2500 platform with 50 bp paired end sequencing. The adapter sequence was AATGATACGGCGACCACCGAGATCTACACTCTTTCCCTACACGACGCTCTTCCGATCT. Data was analyzed essentially as described ^[335]102 . 96% of reads could be aligned against the database Saccharomyces_cerevisiae.EF4.69.dna_sm.toplevel.fa obtained from Ensembl using bwa v0.5.9-r16 ^[336]103 after sequencing adapter removal; duplicates were marked using Picard “MarkDuplicates” and coverage was between 61x and 452x. SNP calling was done after local realignment using “UnifiedGenotyper” from the GATK v2.2 pipeline with default (sensitive) parameters. Annotation was done using ANNOVAR ^[337]104 with information from Ensembl Saccharomyces cerevisiae v69. Positions with coverage below 10x in any of the experiments were excluded from the analysis, as well as positions which were mutated in any of the wild type timepoints as found in > 1 % of reads with a minimum of 2 reads. Suppressor mutations of petites were included if they were found in > 1 % of reads with a minimum of 2 reads at the final time point of at least one replicate culture. Growth curves and chemostat experiments Determination of growth rates For determination of growth rates, yeast was inoculated at an optical density (OD[600]) ~ 0.1 in the respective liquid medium in 96-well plates and cultured in quadruplicates in a FLUOstar OPTIMA plate reader (BMG LABTECH) for 40 h at 30°C, with recording of OD[600] every 20 min. Maximum specific growth rate (μ[max]) and lag phase were both determined from growth curves using a model-richards fit from the R ‘grofit’ package v. 1.1.1-1 ^[338]105 . Chemostat experiments and quantification of fermentation products 200 mL pre-culture medium (per liter: 10 g yeast extract (Merck), 20 g soy peptone (Kerry Bio-Science), 22 g glucose monohydrate (Roth) were inoculated with 750 μL cryostock of the respective yeast strains and grown on an orbital shaker at 30°C and 180 rpm overnight. This culture was used for inoculation of the bioreactors at an OD[600] of 0.4. The batch cultivations were performed in 1.4 L bioreactors (DASGIP Parallel Bioreactor System) with 1.0 L F1 medium containing 2 % glucose. Cultivation temperature was controlled at 30°C, pH was controlled at 5.0, and the dissolved-oxygen concentration was maintained above 20 % saturation by controlling the stirrer speed between 200 to 1200 rpm as well as the air flow between 6 sL/h to 40 sL/h. Inlet and outlet gases were followed with the off-gas sensor provided by the bioreactor system (DASGIP Off-Gas Analyzer GA4, DASGIP AG). Foam formation was prevented by the automatic addition of a 1 % solution of Struktol SB2121 (Schill+Seilacher). 12 mL of samples were taken at regular intervals throughout the whole cultivation duration. Biomass production (OD[600]) was determined using a spectrophotometer (Ultrospec 1100pro, Amersham Biosciences). A correlation between the OD[600] and the cell dry mass (CDM) was established. The concentrations of glucose, glycerol and ethanol in the culture broth were determined by HPLC analysis (Shimadzu Corporation) with a Phenomenex Rezex ROA column (300 mm × 7.8 mm) and a refractive index detector (RID-10A, Shimadzu Corporation). The column was operated at 60°C temperature, 1 mL/min flow rate and 4 mM H[2]SO[4] as a mobile phase. HPLC samples were prepared by adding 100 μL of 40 mM H[2]SO[4] to 900 μL culture supernatant. Subsequently the samples were filtered on 0.20 μm RC membrane filters and 10 μL were injected for analysis. Rates for glucose uptake, ethanol / glycerol production as well as growth rate were estimated by fitting the time courses (including time points up to 10.5 hours of growth) to an ordinary differential equation model (ODE) assuming exponential growth and constant yield. The model was used as described elsewhere ^[339]106 , and gPROMS Model Builder (v. 4.0, Process Systems Enterprise Ltd.) was used for parameter estimation. Microscopy Quantification of mitochondrial membrane potential and morphology by MitoLoc Microscopic analysis of ΔΨ^mito using the MitoLoc system was carried out as previously described ^[340]50 . Pixel-by-pixel colocalization of preSU9-GFP and preCOX4-mCherry was used as a measure for ΔΨ^mito using the Pearson Correlation Coefficient (PCC) as described ^[341]50 . Mitochondrial fragmentation was determined as described ^[342]50 . Quantification of mitochondrial membrane potential with DIOC[6] analyzed by microscopy ΔΨ^mito analysis using 3,3′-dihexyloxacarbocyanine iodide (DiOC[6]) was performed by growing strains in SM for 4 h to an OD[600] of ~0.5. Where indicated, cells were pre-treated with 15 μM CCCP for 5 min to depolarize mitochondria, and 5 mio cells stained with 175 nM DiOC[6] in HEPES with 5 % glucose for 15 min at 30°C. After washing with PBS, cells were resuspended in fresh medium, embedded in agarose pads and subjected to microscopy as described ^[343]50 , using excitation/emission wavelengths of 488/528 nm. Quantification of mitochondrial membrane potential with DIOC[6] analyzed by flow cytometry Freshly streaked colonies were resuspended in 30 μL water and 10 μL of suspension were used to inoculate 500 μL SM or SM with 2mM glutamate, glutamine, arginine, and leucine (+QERL) and incubated overnight with agitation at 30°C. The following day, cultures were diluted 1:2 and cultured for another 4-8 hours to an OD[600] ~0.5. Cells were then spun down at 4000 rpm for 3 min and the media exchanged with SM or SM+QERL supplemented with 175 nM DiOC[6], incubated at 30°C with agitation for a further 15 min. Cells were then repelleted and washed with 500 μL/well PBS, before final centrifugation and resuspension in 500 ul PBS. 200 μL of each culture was transferred to a U-bottomed 96 well plate and taken for high-throughput flow cytometry analysis on a Fortessa X20 HTS cytometer. Acquisition parameters were 120 μL sample injection at a flow rate of 1 μL/sec. DiOC[6] fluorescence was captured using 488 nm blue laser excitation followed by a 530/30 bandpass filter. Quantification of peroxisomes The number of peroxisomes was determined in GFP-Pts1p expressing BY4741 yeast using super-resolution microscopy as described ^[344]50 . In short, BY4741 expressing GFP-Pts1p was cultured to exponential phase in SC-HU medium containing 2 % glucose, harvested by centrifugation, washed twice in PBS and fixed by resuspension in formaldehyde solution (4 g/L PFA, 3.6 % sucrose). The number of peroxisomes was counted using CellProfiler software 2.0 ^[345]107 . ROS quantification ROS production was assessed by growing strains in SM for 4 h to an OD[600] of ~0.5. Where indicated, cells were pre-treated with 2 mM tert-Butyl hydroperoxide (tert-BOOH) for 5 min to induce formation of ROS, and and 5 mio cells stained with 2.5 μg/mL dihydroethidium (DHE) for 5 min in the dark. After washing with PBS, cells were resuspended in fresh medium, embedded in agarose pads and subjected to microscopy as described ^[346]50 , using excitation/emission wavelengths of 510/595 nm. Images were analyzed using CellProfiler software 2.0 ^[347]107 , counting the proportion of cells with detectable fluorescence. Metabolomics and proteomics Metabolomics For quantification quantification of intracellular metabolites, single colonies of respective yeast strains were grown to mid-log phase (OD[600] ~0.75) in selective medium, and a culture volume equalling 7.5 OD[600] was harvested by injecting into cold methanol and subsequent metabolite extraction as described ^[348]73 . Amino acid quantification by LC-MS/MS was performed as previously described ^[349]108 . Sugar phosphate quantification by LC-MS/MS was performed as previously described ^[350]73 . Quantification of ADP, ADP and AMP was performed as described ^[351]109 . For quantification of reduced and oxidized glutathione we performed LC-MS/MS experiments on a 6460 Triple Quadrupole Mass Spectrometer (Agilent) coupled to UPLC (1290 Infinity, Agilent) operated with a binary gradient using 0.1 % formic acid in water as buffer A and methanol as buffer B and a Zorbax Eclipse Plus C18 column (Agilent, RRHD, 2.1x100 mm, 1.8 μm, column temperature: 20°C). Chromatographic separation of extracted analytes was achieved by isocratic flow at 100 % A at 0.5 mL/min for 0.5 min, which was then ramped to 10 % B within 2.5 min. After a 0.5 min washing step at 10 % B the column was re-equilibrated to 100 % A resulting in a total cycle time of 5 min. GSSG and GSH were quantified in positive ESI mode via the MRM transitions 613→484, 613→355, 613→231 (fragmentor: 120, collision energy: 18), and 308→233, 308→179, 308→161 (fragmentor: 105, collision energy: 9), respectively. Medium glucose was quantified using LC-MS/MS as described ^[352]110 modified by including the transition (179→89, 70 V Fragmentor, 5 V collision energy, negative mode). Protein profiling, data analysis, and pathway enrichment For protein profiling, single colonies of respective yeast strains were grown to mid-log phase (OD[600] ~0.7) in selective medium containing 2 % glucose, harvested by centrifugation and snap-frozen in aliquots equalling 10 OD[600] units. Sample preparation was carried out as previously described ^[353]60 . SWATH LC-MS/MS analysis was performed essentially as previously described ^[354]60 on a TripleTOF5600 instrument (SCIEX) online coupled to a nanoACQUITY chromatographic system (Waters) operating at 3 μL/min flow rate. Data was analyzed with Spectronaut software (version 14, Biognosys AG) and post-processed in statistical language R. Principal component analysis was carried out using the unfiltered data set and a prcomp function. Protein fold change was calculated with reference to the wild type strain, and differential abundance was defined as a fold change > 1.5 and FDR-corrected p-value < 0.01. Pathway enrichment was performed in String-db v11 ^[355]111 , and Reactome pathways ^[356]112 were reported as significantly enriched if FDR-corrected enrichment p-value < 0.05. Dynamic labelling experiments Sample preparation For dynamic labelling experiments, yeast cultures were inoculated at OD[600] of 0.15 in F1 medium containing 2 % glucose and agitated at 30°C until reaching mid-log phase. 45 OD[600] units (~6 × 10^8 cells) were pelleted, re-suspended in 6 mL F1 medium containing 1 % (w/v) [U-^13C[6]]-glucose (Cambridge Isotope Laboratories) and kept at 30°C with occasional agitation. 1 mL of cells (~1 × 10^8 cells) was removed after 1, 5, 15 and 30 min, shock-frozen by injection into 20 mL of -40°C MeOH and pelleted by centrifugation (2 min, 4000 × g, 4°C). Methanol was removed, and metabolites extracted by resuspension of the pellet in 600 μL chloroform/methanol (600 μL, 2:1 v/v), transfer to a 1.5 mL tube, brief vortexing, and subsequent periodic sonication in a water bath for 1 hr at 4°C. Samples were pelleted by centrifugation (10 min, 16000 ×g, 4°C), the metabolite extract transferred to a new tube, and the pellet re-extracted with methanol/water (600 μL, 2:1 v/v, containing 1 nmol scyllo-inositol). Samples were pelleted as above, the second extract added to the first and dried in a rotary vacuum concentrator. Samples were biphasic partitioned by addition of 100 μL chloroform, followed by 600 μL methanol/water (1:1 v/v) and vigorous vortexing. After centrifugation (as above), the upper (polar) phase was removed, dried, and processed for gas chromatography-mass spectrometry (see below). Gas Chromatography-Mass Spectrometry (GC-MS) Polar metabolites were derivatized and analysed by GC-MS (Agilent 7890B-5977A) and identification and abundance of individual metabolites was estimated as previously described ^[357]113 . In brief, dried metabolite samples were washed with methanol (twice), and derivatized overnight at RT with methoxyamine (20 mg/mL in pyridine, Sigma) followed by addition of BSTFA + 1% TMCS (Sigma) for > 1 hr at RT. GC-MS was performed using splitless injection (injection temperature 270°C) onto a 30 m + 10 m × 0.25 mm DB-5MS+DG column (Agilent J&W), with helium carrier gas, in electron impact ionization mode. The oven temperature was initially 70°C (2 min), followed by a temperature increase to 295°C at 12.5°C/min and subsequently to 320°C at 25°C/min (held for 3 min). GAVIN software ^[358]114 was used for metabolite identification and quantification by comparison to the retention times, mass spectra, and responses of known amounts of authentic standards. Iron depletion experiments, and ^55Fe incorporation binding assays Iron depletion for proteomics Culture was performed by inoculating a pre-culture (iron-free SM medium with or without iron, 2 % glucose) with washed yeast cells and incubation overnight at 30°C. Main culture was inoculated to OD[600] of 0.1, and incubated at 30°C with shaking until cultures reached OD[600] of 0.7. 10 OD[600] units were collected by centrifugation, washed with iron-free water and stored at -80°C. Iron depletion in liquid media Single colonies of indicated strains were transferred to 10 mL SM situated in Erlenmeyer flasks, and incubated overnight. Cultures were diluted to OD[600] ~0.2 and incubated for 5 hours. Cells were harvested, supernatants were discarded, and cells were washed 3 times in sterile MiliQ H[2]O, and OD[600] was adjusted to 0.2. Then, 5 μL of each culture was transferred to a microtiter plate with 195 μL of minimal media with iron, without iron, and without iron and increasing dipyridyl (DIP) concentration. The base formulation of this media corresponds to 6.7 g/L yeast nitrogen base (YNB, Sigma Y0626) with 2 % glucose. For iron depletion, the ferric chloride was omitted, and where indicated DIP was added from a 50 mM stock dissolved in DMSO to a final concentration of 50, 100, 150, 175 μM respectively. Growth curves were obtained on a Tecan Spark-Stacker using the mean values of five multi-well reads at OD[600] obtained every 30 minutes. Growth rates were estimated from obtained growth data using the spline fit implemented in the R package grofit (v1.1.1-1). ^55Fe incorporation assay Quantification of Fe-S cluster formation was performed essentially as described ^[359]115 . BY4741 yeast transformed with pUG35_Aco1 construct were verified for mitochondrial GFP expression using fluorescence microscopy, and pre-cultured overnight in iron-free SM+LMW medium with 2 % glucose. Main culture was inoculated to OD[600] ~0.1, and incubated at 30°C until all strains were in mid-log phase. Cells were harvested by centrifugation, washed once with analytical grade water, re-suspended in 10 mL fresh medium per 0.5 g wet cell mass, and incubated for 10 min (30°C, 200 rpm). 10 μCi ^55FeCl[3] (Perkin Elmer) in 100 μL 0.1 M sodium ascorbate was added to each sample, and iron incorporation performed for 2 h. After collecting cells by centrifugation (3000 xg, 5 min), pellet was washed once with 10 mL citrate buffer (50 mM sodium citrate, 1 mM EDTA, pH 7.0) and once with 1 mL HEPES-KOH pH 7.4. Cells were disintegrated by bead shaking in a FastPrep device (MP Biomedicals) with 500 μL TNETG+PI buffer (20 mM Tris-HCl pH 7.4, 2.5 mM EDTA, 150 mM NaCl, 10 % (w/v) glycerol, 0.5 % (w/v) Triton X-100, 1x cOmplete Protease Inhibitor Cocktail (Roche) and 0.5 volume acid-washed glass beads for 30 s at 25/s speed. After centrifugation (10 min, 17000 xg, 4°C), protein concentration of supernatants was determined, adjusted to 15 μg/μL and GFP-Aco1p enriched using 25 μL washed GFP-Trap magnetic beads (ChromoTek). Samples were inverted (1 h, 4°C), beads collected and washed extensively with TNETG+PI buffer. Bound ^55FeCl[3] was quantified by scintillation counting. ICP-MS For quantification of elemental composition, yeast cells were resuspended in 4.5 mL analytical grade water, and proteins were cleaved by addition of 1.4 mL concentrated HNO3 (final concentration 16.4 %) and incubation for 2 h at 80°C. Following centrifugation (10 min, 4000 xg), the supernatants were then analyzed undiluted in triplicate using inductively coupled plasma spectrometry (ICP-MS) on a Perkin Elmer Elan DRC II instrument. The instrument was calibrated by dilution of Fe standard solution (1 ppb to 1 ppm) and the standard was re-analyzed every 16 samples to correct for instrumental drift. A 1 ppb solution of In, Re and Rh was used as an internal standard. Concentrations of elements were normalized to protein concentration of yeast lysate as determined by BCA assay. Biochemical and molecular biology assays Rapamycin sensitivity testing For rapamycin sensitivity assay, strains were inoculated at 0.1 OD[600] with 11 rapamycin concentrations from 50 pg/mL to 1 μg/mL (n = 3) in a 384 multi-well plate (70 μL/well) in SC-H medium and growth was monitored for 48 h. The maximum specific growth rate and EC50 values for dose response curves were determined with R package grofit (v1.1.1-1). Yeast dilution spot tests For oxidant spot tests, a 2 mL overnight culture of yeast was prepared. Cells were harvested, washed once in H2O and re-suspended in fresh, autoclaved H2O to an OD[600] of 2.5 unless otherwise indicated. 200 μL of the resulting suspension was transferred to a 96-well plate and a 1:5 dilution series was prepared. Oxidant plates were prepared freshly by cooling media containing agar to ~50°C, 1.2 mM diamide was added to aliquots, mixed, cast into petri dishes and cooled to RT. Cell suspensions were vigorously mixed, and 5 μL of each dilution was spotted onto agar plates. Spots were air-dried under a flow hood and incubated at 30°C for 2-3 days. For amino acid supplementation spot tests, overnight cultures of indicated strains were prepared in 15 mL SM situated in flasks, and supplemented with or without indicated amino acids. The overnight culture was diluted 1:20 in respective media, and incubated for 5 hours. Cells were harvested, set to an OD[600] of 0.5 in the respective media, and spotted and incubated as described above. For iron depletion spot tests, the indicated strains were pre-grown in SM with 2 % glucose. A serial dilution series (1:5) was spotted with an initial OD[600] of 0.5 on SM agar plates with or without 1 mM ferrozine. Quantification of enzyme activities Isocitrate dehydrogenase activity was quantified using a coupled diaphorase assay as described ^[360]116 , using 20 μg/mL yeast lysate and 2 mM NAD^+ or 2 mM NADP^+. The reaction was started by addition of 16 mM isocitrate and followed continuously in 96-well plates at 340 nm in a plate reader (Tecan Infinite 200Pro). Aconitase activity was quantified as described ^[361]99 , with 1 U/mL IDH1 (Roche), 0.4 mM NADP^+, and 20 μg/mL yeast lysate. The reaction was started by addition of 2 mM citrate and absorption at 340 nm followed continuously as above. ATPase (complex V) activity was quantified as described ^[362]117 , and specific activity (μmol ADP/mg protein/min) of the oligomycin-insensitive fraction was calculated according to the Beer-Lambert law equation. Aconitase overexpression and localization tagging Wild type and petite BY4741 was transformed with a plasmid expression of Aco1p (ACO1), Aco1p constitutively targeted to mitochondria (mtACO1), or the empty vector. Cells were grown to mid-exponential phase in SC-U medium, harvested, broken, and aconitase activities determined as described above. Supplementary Material Suppl. Fig. 1 [363]EMS140507-supplement-Suppl__Fig__1.pdf^ (3.6MB, pdf) Suppl. Table 1 [364]EMS140507-supplement-Suppl__Table_1.xlsx^ (1.4MB, xlsx) Acknowledgements