Abstract
The human ATP synthase is an assembly of 29 subunits of 18 different types, of which only two (a and 8) are encoded in the mitochondrial genome. Subunit a, together with an oligomeric ring of c-subunit (c-ring), forms the proton pathway responsible for the transport of protons through the mitochondrial inner membrane, coupled to rotation of the c-ring and ATP synthesis. Neuromuscular diseases have been associated to a number of mutations in the gene encoding subunit a, ATP6. The most common, m.8993 T > G, leads to replacement of a strictly conserved leucine residue with arginine (aL156R). We previously showed that the equivalent mutation (aL173R) dramatically compromises respiratory growth of Saccharomyces cerevisiae and causes a 90% drop in the rate of mitochondrial ATP synthesis. Here, we isolated revertants from the aL173R strain that show improved respiratory growth. Four first-site reversions at codon 173 (aL173M, aL173S, aL173K and aL173W) and five second-site reversions at another codon (aR169M, aR169S, aA170P, aA170G and aI216S) were identified. Based on the atomic structures of yeast ATP synthase and the biochemical properties of the revertant strains, we propose that the aL173R mutation is responsible for unfavorable electrostatic interactions that prevent the release of protons from the c-ring into a channel from which protons move from the c-ring to the mitochondrial matrix. The results provide further evidence that yeast aL173 (and thus human aL156) optimizes the exit of protons from ATP synthase, but is not essential despite its strict evolutionary conservation.
Introduction
Mitochondrial disorders result from defects in oxidative phosphorylation (OXPHOS) (1,2). Oxidative phosphorylation uses the energy obtained from the catabolism of the food we ingest for the synthesis of adenosine triphosphate (ATP), which is the main energy source for the cellular processes that sustain life (3). OXPHOS typically involves five multiprotein complexes (I–V) all of which are embedded or anchored in the inner membrane of mitochondria. Complexes I–IV, which constitute the electron transport chain (ETC), obtain electrons from the oxidation of carbohydrates and fatty acids, which are ultimately used to reduce oxygen to water. In the process of oxidation/reduction reactions in the ETC, protons are pumped out of the matrix to the intermembrane space (IMS) establishing a proton motive force across the mitochondrial membrane. The proton motive force is used by the ATP synthase, Complex V, which couples the return of protons into the matrix with phosphorylation of ADP forming ATP (4). Any defect in the activity or biogenesis of the OXPHOS system will affect high-energy demanding tissues and organs and especially the brain, which consumes two-thirds of the dozens of kilograms of ATP that our body uses each day.
The human mitochondrial genome encodes 13 of the 90 subunits of the OXPHOS system, and a number of RNAs required for their synthesis inside the mitochondria. The remaining subunits are encoded in the nuclear genome and post-translationally imported into the mitochondrion. Despite the preponderance of nuclear genes required for formation and functioning of the OXPHOS system, mutations in the mitochondrial genome have been most frequently identified in mitochondrial disorders, presumably because of the exposure to this genome to damaging reactive oxygen species (ROS) (5) and the poor effectiveness of the mitochondrial DNA (mtDNA) repair systems (6). An added complexity for diseases due to mutations in the mtDNA is that the mitochondrial genome is present in up to 10 000 copies in a single cell (7,8). Thus, pathogenic mutations in the mtDNA usually co-exist in patient’s cells and tissues along with wild-type copies of mtDNA (heteroplasmy). This genetic heterogeneity can make it difficult to understand precisely how and to what extent these mutations compromise mitochondrial function and thus human health.
Yeast Saccharomyces cerevisiae is an unicellular fungus amenable to manipulation of the mitochondrial genome (9) and is unable to stably maintain mitochondrial heteroplasmy (10). Owing to its efficient capacity to produce ATP by the fermentation of sugars, this yeast can survive mutations that inactivate oxidative phosphorylation. We have exploited these attributes to investigate the consequences of one of the most common mtDNA mutation (m.8993 T > G) responsible for human mitochondrial diseases, including neuropathy ataxia retinitis pigmentosa (NARP) and maternally inherited leigh syndrome (MILS) (11–23). The m.8993 T > G mutation is in the gene encoding subunit a of the ATP synthase (ATP6) and changes a strictly conserved leucine residue to arginine, aL156R. Subunit a and an oligomeric ring of identical c-subunits (c-ring) are responsible for the transport of protons from cytoplasm to the matrix coupled to the synthesis of ATP (24–27). Previously, we showed that an equivalent of the m.8993 T > G mutation in yeast (aL173R) severely compromised (by 90%) mitochondrial ATP synthesis (28). Here, we report nine intragenic pseudo-reversions of this mutation that restore to various degrees ATP synthase function. Considering the recently described high-resolution structures of yeast ATP synthase (25,29), these findings suggest the molecular bases of the primary pathogenic effects of the m.8993 T > G mutation and provide novel information on the proton pathway within the membrane domain of ATP synthase.
Results
Selection of genetic suppressors of the aL173R mutation
Genetic reversion of the mutant phenotype induced by aL173R can result from a second mutation within the ATP6 gene (intragenic suppression) or in a different gene (extragenic suppression) located in mtDNA (intragenomic suppression) or nuclear (intergenomic suppression) DNA. Owing to the inability of S. cerevisiae to maintain mitochondrial heteroplasmy, genetic suppressors of aL173R in mtDNA will be present in all mtDNA molecules (homoplasmy). With a haploid strain (in which nuclear genes are in single copy), nuclear suppressors can be dominant or recessive, whereas only dominant suppressors can be selected from a diploid strain where each nuclear gene is in two copies. We, therefore, used in this study diploid (MR12) and haploid (MR14) yeast strains for the isolation of genetic suppressors of the aL173R mutation.
From diploid strain MR12
The m.8993 T > G mutation converts leucine at position 156 of human subunit a with arginine (aL156R). The corresponding residue in the mature form of yeast subunit a is aL173 (aL183 in the subunit a polypeptide precursor from which the first 10 N-terminal residues are removed during its maturation (Fig. 4A)) (30). In the mutant diploid strain, MR12, the corresponding leucine codon TTA was replaced with the arginine codon AGA (aL173R, TTA183AGA) (see Materials and Methods). As expected from the severe consequences of the aL173R change on ATP synthase (28), MR12 grew very poorly on non-fermentable substrates as compared with the parent strain, MR11 (Fig. 1A).
Figure 4.

Molecular basis of the detriment effect of aL173R. (A) Sequence alignment of parts of aH5 (157–191) (right) and aH6 (a.a. 215–245) of subunits a (left) from various origins The shown sequences are from Saccharomyces cerevisiae (S.c.), Yarrowia lipolytica (Y.l.), Schizosaccharomyces pombe (S.p.), Polytomella sp (P.sp.), Arabidopsis thaliana (A.t.), Bos taurus (B.t.), Sus scrofa (S.s.) and Homo sapiens (H.s.). Positions of first-site and second-site reversions in haploid and diploid strains are indicated. (B) Overall top view from the matrix of the a/c10-ring with the pathway along which protons move from the intermembrane space (IMS) to the mitochondrial matrix. The side chains of residues (depicted with balls and sticks) and the n-side (purple) and p-side (green) pockets involved in this transfer are shown. Positions of first-site and second-site reversions and crucial residues for proton translocation are indicated. (C) View from the entry-to-exit channels of amino acid residues surrounding aL173. (D) View of the subunit a portion at the interface with the c-ring are drawn as balls and sticks side chains of highly conserved and important residues. Residue side chains coloured in green on a pale green background form an hydrophobic barrier between the two hydrophilic clefts; those strictly conserved are underlined. The entry and exit gates for protons to and from the c-ring are encircled, and their routes within the two half-channels of subunit a are depicted by the green and purple arrowhead lines. (E) View in the same orientation as (C) of the indole nitrogen of aW173 presumably responsible for a partial proton shortage between to the hydrophilic clefts in the aL173W strain. (F) While the aliphatic part of aR173 contributes to the hydrophobic sealing of the n- and p-side clefts, its guanidinium group prevents the exit of protons from the c-ring in proximity of aY241. (G) The position of the guanidium group of aK173 and its hydrophobic alkyl chain enable proton release from the c-ring in the p-side cleft. (H) Replacing aR169 by S or M at the proton exit gate possibly shift and partially restore proton flow despite the presence of aR173. The residue color code for all the panels is green for hydrophobicity, cyan and magenta for entry and exit channels, respectively, and yellow for second-site reversions of aL173R.
Figure 1.

Influence of the subunit a mutations on yeast respiratory growth. (A) Fresh glucose cultures of the diploid (MR12) and haploid (MR14) aL173R mutants, their revertants (designated by their mutation(s) in subunit a) and parental strains (MR11 and MR6, respectively) were serially diluted and spotted onto rich glucose and glycerol plates. The shown plates were scanned after 4 days incubation at 28°C. (B) Growth of the same strains in liquid glycerol medium monitored with the bioscreen system over a period of 80 h during which cell densities (OD600nm) were taken every 20 min.
Mutant derivatives of MR12 that suppressed or reverted the disease mutation were selected from 22 colonies of MR12 individually grown in glucose media. Cells (108) of each culture were spread on an agar plate containing glycerol as the sole carbon source. On average, 10 colonies grew on each selection plate, from which only 1 or 2 were taken for further analysis. In this way, most of the retained isolates were genetically independent (i.e. not daughter cells). These strains all grew significantly better on medium containing glycerol as the carbon source as compared with the original aL173R mutant strain (Fig. 1A). Some of the revertants grew comparable to the wild-type yeast whereas others grew much less efficiently. Because of the nature of the initial mutation, none of the isolates reverted the AGA codon to initial leucine TTA codon as this would have required an extremely rare (about 10−14) double nucleotide change (AGA183TTA), and there was no possibility to revert to another leucine codon by a single nucleotide substitution. Sixteen revertants had a mutation that altered the TTA codon into a different amino acid codon (first-site reversion) with AGA183ATA (aR173M) occurring in 12 clones, AGA183AAA (aR173K) in 3 clones and AGA183AGT (aR173S) in 1 clone (Table 1). In the six remaining revertants, a single nucleotide change was identified in another ATP6 codon (second-site reversion): AGA179AGT (aL173R + aR169S) in 3 clones, GCT180CCT (aL173R + aA170P) in 1 clone, GCT180GGT (aL173R + aA170G) in 1 clone and ATC226AGC (aL173R + aI216S) in 1 clone. Upon sporulation, all the revertants produced asci in which all four spores had a better respiratory growth than spores issued from MR12 (aL173R) indicating that a nuclear mutation was not contributing to the improved growth phenotype. For the sake of clarity, the first-site suppressors (aR173M, aR173K and aR173S) will be identified relative to the wild-type subunit a sequence (aL173M, aL173K and aL173S).
Table 1.
Intragenic suppressors of aL173R
| Codon change |
a.a change |
From MR12 | From MR14 |
|---|---|---|---|
| First-site reversions | |||
| AGA183ATA | 𝑎R173M | 12 | 32 |
| AGA183AAA | 𝑎R173K | 3 | 5 |
| AGA183AGT | 𝑎R173S | 1 | 0 |
| AGA183TGA | 𝑎R173W | 0 | 1 |
| Second-site reversions | |||
| AGA179AGT | 𝑎R169S | 3 | 0 |
| AGA179ATA | 𝑎R169M | 0 | 1 |
| GCT180CCT | 𝑎A170P | 1 | 0 |
| GCT180GGT | 𝑎A170G | 1 | 0 |
| ATC226AGC | 𝑎I216S | 1 | 0 |
From haploid strain MR14
Using the same procedure used for the isolation of revertants from MR12, we isolated revertants from the previously described aL173R haploid strain MR14 (MR6 is the parent), which had the same mutant codon (TTA183AGA) (28). We identified four different intragenic pseudo-reversions: AGA183ATA (aR173M) in 32 clones; AGA183AAA (aR173K) in 5 clones; AGA183TGA (aR173W) in one clone and AGA179ATA (aL173R + aR169M) in one clone. While the three first-site reversions conferred a good growth on glycerol medium, the strain with the second-site reversion did not grow as well on non-fermentable carbon sources (Fig. 1B).
As expected, the revertant mutations isolated from MR12 and MR14 overlapped. The effect of the mutation aL173R in both strains was indeed suppressed with replacement of L173 with Lys or Met, and the same second-site suppressor site was identified in codon for aR169 and the resultant growth phenotypes were comparable.
Influence on the OXPHOS system of the genetic suppressors of aL173R (isolated from strain MR14)
During the course of this study, we determined that MR11, and thus, all its derivatives (MR12 and revertants), contained an additional mutation (cL53F) in the mitochondrial gene encoding subunit c (ATP9), which was determined to substantially compromise ATP synthase function (c.f. Supplementary Material, Info, Figs S1–5). We, therefore, investigated the functional influence of the genetic suppressors only isolated from strain MR14, which did not have the subunit c mutation. However, as cL53 is far (12.4 Å) from aL173 in yeast ATP synthase structure (Supplementary Material, Fig. S3C), and because of the overlap between the suppressors identified from MR12 and MR14 (Table 1), we are confident that those isolated from MR12 correct the detrimental consequences of aL173R even in the presence of cL53F (see below).
Assembly of OXPHOS complexes
The influence on the assembly/stability of the four genetic suppressors of aL173R (aR173W, aR173K, aR173M, and aL173R + aR169M) selected from strain MR14 was investigated by BN-PAGE of mitochondrial samples solubilized with digitonin (Fig. 2A–E), and by evaluation of the steady state levels of subunit a by western blot analysis (Fig. 2G). ATP synthase complexes were visualized in BN gels by ATPase activity (Fig. 2A), Coomassie blue staining (Fig. 2B), and by using antibodies against subunit c (Atp9) (Fig. 2C), subunit a (Atp6) (Fig. 2D) and the α subunit of F1 (Atp1) (Fig. 2E). In the conditions used, ATP synthase monomers were much more abundant than oligomers of ATP synthase. There is evidence that most, if not all, of the ATP synthase subunits can assemble in the absence of subunit a, as was observed in yeast strains lacking the ATP6 gene or that of a protein, Atp10, specifically required to assemble subunit a (31,32). However, the resulting assemblies dissociate easily as evidenced by the presence of free F1 and c-ring in BN-gels. As was already observed (28), the ATP synthase accumulated less efficiently in the aL173R strain compared with WT yeast, but free F1 and c-ring were not detected indicating that the mutated subunit a assembles normally. Since unassembled yeast subunit a is rapidly degraded (33,34) the steady state levels of subunit a in denaturing gels provide better estimates of the levels of fully assembled ATP synthase than BN-PAGE. Based on the level of subunit a, the amount of holo-ATP synthase in aL173R samples was at 58% versus WT (Fig. 2G). Since mitochondrial biogenesis is an energy-consuming process, the difference was likely the consequence of the extremely low ATP producing capacity of the aL173R mitochondria (10% versus WT (31), see below), rather than specific defects in the assembly of the ATP synthase. Also, partial ATP synthase intermediates were not detected in the samples of the revertants. Good accumulation of subunit a was observed in mitochondria from the aR173W, aR173K, and aR173M strains (77–91% versus WT), whereas the aL173R + aR169M strain (54%) had a deficit in subunit a similar to that observed in the aL173R mutant (58%). The amounts of ATP synthase monomers in BN-gels shown in Figure 2C–E were rather consistent with the levels of subunit a in denaturing gels (Fig. 2F).
Figure 2.

Influence of the subunit a mutations on ATP synthase assembly and accumulation levels. Mitochondria (250 μg) isolated from wild-type yeast MR6, the aL173R strain MR14 and its revertants (designated by their mutation(s) in subunit a) were solubilized with digitonin (2 g/g protein) and separated by BN-PAGE in a 3–12% polyacrylamide gel. ATP synthase complexes (monomer, V1 and oligomers, Vn) were in-gel detected with their ATPase activity (A), Coomassie blue (B), and with antibodies against subunits Atp9/c (C), Atp6/a (D) and Atp1/αF1 (E) after their transfer to a PVDF membrane. (F) The amounts of ATP synthase monomers, expressed in %WT and normalized to the content of porin in the samples, were estimated from Fig. 2C-E (with ImageJ). (G) Mitochondrial proteins were separated by SDS-PAGE (10 μg per lane), transferred to a nitrocellulose membrane, and probed with antibodies against the indicated proteins. The strains corresponding to lanes 1–6 are indicated. The shown protein detection assays are representative of at least three independent experiments.
We further evaluated the abundance of cytochrome oxidase (Complex IV) by measuring the steady levels of the Cox2 subunit of this complex this protein is degraded when unassembled (35). Secondarily, yeast ATP synthase defective mutants also have a reduced content in cytochrome oxidase except for those that have a large impact on proton leaks through FO (28,31,36–39), indicating that the biogenesis of this complex is modulated by the activity of FO (40). Mitochondria from the aL173R and the aL173R + aR169M strains showed a large drop in the content of cytochrome oxidase (80% versus wild type), whereas it was much more abundant in the aR173W, aR173K and aR173M strains (Fig. 2F).
Mitochondrial respiration and ATP synthesis
We next investigated the influence on respiration (oxygen consumption) and ATP synthesis of the genetic suppressors of aL173R, in intact (osmotically protected) mitochondria. Oxygen consumption was measured with NADH and ascorbate/TMPD as sources of electrons. The transfer of electrons to oxygen from NADH involves the NADH dehydrogenase located at the external surface of mitochondria (Nde1) and cytochrome b-c1 reductase and cytochrome oxidase, whereas ascorbate/TMPD delivers electrons directly to cytochrome oxidase. NADH was used alone (State 4, where respiration is induced only by the passive permeability to protons of the inner membrane), in the presence of ADP (State 3 or phosphorylating conditions) or the uncoupler agent CCCP (thus without any membrane potential, conditions where respiration is maximal). The rate of ATP synthesis was measured in the presence of NADH and a large excess of external ADP.
As already reported (28), electrons from NADH and ascorbate/TMPD were transferred slowly to oxygen in aL173R mitochondria (20% versus WT, see Table 2), because of their low content in cytochrome oxidase (see above), and the rate of ATP synthesis was reduced in similar proportions (Table 2). These activities were minutely increased by the second-site suppressor of aL173R (aR169M), which is in line with its very modest respiratory growth improvement versus aL173R (Fig. 1B). On the other hand, consistent with their good growth on glycerol medium (Fig. 1), the levels of respiration and ATP synthesis were considerably improved in the first-site revertants (aR173W, aR173K, aR173M) (Table 2).
Table 2.
Mitochondrial respiration and ATP synthesis
| Strain | Respiration nmol O. min−1.mg−1 (% WT in parenthesis) |
ATP synthesis nmol Pi. min−1.mg−1 |
||||||
|---|---|---|---|---|---|---|---|---|
| NADH | NADH+ADP | NADH+CCCP | Asc/TMPD +CCCP |
-Oligo | +Oligo | F1FO % WTa |
F1FO % WTb |
|
| WT | 329 ± 26 (100) | 650 ± 73 (100) | 1550 ± 96 (100) | 2809 ± 299 (100) | 640 ± 100 | 15 ± 3 | 100 | 100 |
| 𝑎L173R | 137 ± 17 (42) | 132 ± 32 (20) | 279 ± 38 (18) | 489 ± 66 (17) | 48 ± 13 | 10 ± 3 | <10% | 10 |
| 𝑎L173W | 341 ± 14 (104) | 675 ± 43 (104) | 1558 ± 78 (101) | 2548 ± 80 (91) | 488 ± 27 | 16 ± 1 | 76 | 83 |
| 𝑎L173K | 265 ± 24 (81) | 422 ± 27 (65) | 986 ± 69 (64) | 1400 ± 34 (50) | 326 ± 36 | 16 ± 2 | 50 | 64 |
| 𝑎L173M | 247 ± 14 (75) | 436 ± 30 (67) | 1052 ± 80 (68) | 1327 ± 208 (47) | 405 ± 26 | 17 ± 4 | 62 | 75 |
| 𝑎R169M | 206 ± 12 (63) | 188 ± 28 (29) | 342 ± 39 (22) | 526 ± 77 (19) | 75 ± 15 | 22 ± 5 | <10% | 16 |
Mitochondria were isolated from cells grown for 5–6 generations in rich galactose medium (YPGalA) at 28 °C. Reaction mixes for assays contained 0.15 mg/mL protein, 4 mM NADH, 150 μM ADP (for respiration assays), 1 mM ADP (for ATP synthesis assays), 12.5 mM ascorbate, 1.4 mM N,N,N,N,-tetramethyl-p-phenylenediamine (Asc/TMPD), 4 μM CCCP, 4 μg/mL oligomycin (oligo). F1FO represents the specific ATP synthesis by F1F0-ATP synthase relative to WT, per mg of protein (a) and after normalization to the relative contents of subunit a (b) determined in Figure 2F.
Mitochondrial ATP hydrolysis
We tested ATP hydrolysis in non-osmotically-protected mitochondria (this to relax the enzyme from any membrane potential that would limit its ATP hydrolytic activity). ATPase was measured at pH 8.4 to prevent binding of the natural peptide inhibitor IF1 to F1 (41). Oligomycin inhibits the ATPase activity of fully assembled and coupled mitochondrial ATP synthase, but not, for instance, free F1-ATPase. In samples with the wild-type enzyme, oligomycin inhibits ATPase activity by about 80% (Table 3). Any oligomycin-insensitive activity is normally due to other ATPases in the preparation. The mitochondrial samples from the aL173R mutant and three of its revertants (aL173W, aL173M, aL173K) showed good oligomycin-sensitive ATPase activity. Surprisingly, while ATP synthesis activity from aL173R was severely diminished, the ATPase activity was not decreased and ATPase was sensitive to oligomycin. This result seems to separate ATPase activity from ATP synthesis activity. In contrast, the ATPase activity of mitochondria isolated from yeast second-site suppressor strain, aL173R + aR169M, was just half the rate relative to the wild type, and it was only inhibited 25% by oligomycin. Thus, while the second-site suppressor mutation aR169M decreased the ATPase activity and oligomycin sensitivity, it increased the rate of ATP synthesis relative to the mitochondria isolated from the strain with aL173R in an otherwise wild-type background.
Table 3.
Mitochondrial ATPase activity
| Strain | ATPase activity μmol Pi. min−1.mg−1 | ||
|---|---|---|---|
| -Oligo | +Oligo | % Inhib. | |
| WT | 6.851 ± 0.292 | 1.623 ± 0.179 | 76 |
| 𝑎L173R | 4.767 ± 0.321 | 1.517 ± 0.496 | 68 |
| 𝑎L173W | 6.881 ± 0.058 | 1.092 ± 0.079 | 84 |
| 𝑎L173K | 6.509 ± 0.042 | 1.405 ± 0.321 | 78 |
| 𝑎L173M | 7.064 ± 0.042 | 1.329 ± 0.213 | 81 |
| 𝑎R169M | 3.832 ± 0.267 | 2.826 ± 0.346 | 26 |
The ATPase assays were performed from mitochondrial samples that had been frozen at −80°C in the absence of osmotic protection at pH 8.4, in the ATPase buffer (see section Materials and Methods).
Mitochondrial membrane potential
We further investigated (using intact mitochondria buffered at pH 6.8) the effect of the suppressor amino acid changes had on the ability of the mitochondria to generate a transmembrane potential gradient (ΔΨ). We assessed this using Rhodamine 123, a cationic dye whose fluorescence is quenched relative to the ΔΨ. The ΔΨ was made by the electron transport chain by oxidation of NADH formed by the oxidation of ethanol. The resulting potential gradient can be diminished by the ATP synthase with the phosphorylation of ADP (Fig. 3A). The addition of ADP normally results in a transient fluorescence increase that progressively disappears until all of the ADP is phosphorylated (Fig. 3A). Thus, this assay evaluates how tightly proton transport is coupled with ATP synthesis and the ability of the ATP synthase to phosphorylate ADP. It is to be noted that the activity of the electron transport chain needs to be strongly diminished (by at least 80%) to observe a reduced ability to create a potential gradient with the oxidation of ethanol. Accordingly, consistent with their respective rates of oxygen consumption (Table 2), mitochondria from the aL173R mutant responded poorly to the addition of ethanol whereas those from the other tested strains generated a higher ΔΨ (Table 2). Consistent with the respective rates of ATP synthesis (Table 2), the potential induced by ethanol in mitochondria from the aL173R and aL173R + aR169M strains did not change when ADP was added, whereas it did so in those from the first-site revertants (aL173W, aL173M, aL173K) (Fig. 3A). This indicates that while the ATP synthase from aL173R can hydrolyze ATP and ATPase is sensitive to oligomycin, it cannot establish a proton gradient with ATP hydrolysis. KCN inhibits the electron transport chain by inhibiting cytochrome oxidase. Normally, KCN addition after ATP synthesis results in a partial loss of membrane potential because the newly formed ATP is hydrolyzed by the ATP synthase, which pumps protons outside the mitochondrion (as evidenced by the loss of the remaining ΔΨ after a further addition of oligomycin (Fig. 3A)). Mitochondria isolated from the strains with the first-site revertants clearly showed this biphasic loss of ΔΨ, whereas the membrane potential was largely and rapidly lost after KCN addition in mitochondria from the aL173R and aL173R + aR169M strains, which further reflects the differences in the activity of the ATP synthase.
Figure 3.

Influence of the subunit a mutations on ATP synthase-dependent variations in the electrical potential across the mitochondrial inner membrane. Variations in mitochondrial transmembrane potential (ΔΨ) were monitored by fluorescence quenching of Rhodamine 123 in intact mitochondria from wild-type yeast MR6, the aL173R strain MR14 and its revertants (designated by their mutations(s) in subunit a as indicated). (A) Tests of ΔΨ consumption by the ATP synthase upon addition of ADP to mitochondria energized with electrons from ethanol. (B) ATP-driven proton pumping assays. The additions were 25 μg/mL rhodamine 123, 150 μg/mL mitochondrial proteins (Mito), 10 μL ethanol (EtOH), 75 μM ADP, 2 mM potassium cyanide (KCN), 0.2 mM ATP, 4 μg/mL oligomycin (oligo) and 4 μM carbonyl cyanide-m-chlorophenyl hydrazone (CCCP). The shown fluorescence traces are representative of at least three independent experiments.
We also examined ATP-driven proton-pumping activity of ATP synthase in the presence of large excess of externally added ATP (Fig. 3B). To this end, mitochondria were energized with ethanol and the resulting ΔΨ was collapsed by adding KCN, releasing inhibitor IF1 from the ATP synthase. ATP was then added less than 1 min later, before IF1 rebinding (42) to determine if ATP hydrolysis will generate a ΔΨ. As expected, a large oligomycin-sensitive ΔΨ variation was observed in mitochondria isolated from the wild-type strain. This was also observed in mitochondria from the first-site revertant strains (aL173W, aL173M, aL173K), but a much reduced proton-pumping activity was observed in mitochondria from aL173R and aL173R + aR169M strains (Fig. 3B). Thus, again, while the ATP synthase of the aL173R mutant showed a good oligomycin-sensitive ATP hydrolytic activity in the absence of any membrane potential (Table 2), it fails to work in the reverse mode in intact mitochondria. Possibly, the aL173R mutation prevents the any significant movement of protons through the FO against a proton gradient.
Structural modeling of the subunit a mutations
Subunits a and c are major components of Fo, which are directly responsible for the movement of protons coupled to the synthesis of ATP by the ATP synthase. Starting from the N-terminus, yeast subunit a has a membrane spanning α-helix (aH1) followed by a 4-helix bundle (aH3-aH6) horizontally wrapped around the c-ring rotor (Fig. 4B) (25). The two domains are linked on the matrix side of the membrane by an amphipathic α-helix (aH2). Two charged residues (aR176 in aH5 and cE59 in cTM2) facing each other near the middle of the membrane at the a/c-ring interface are essential for activity (25,26). Two hydrophilic pockets enable protons from the p-side of the membrane (intermembrane space) to reach cE59 and be released after an almost complete rotation of the c-ring on the n-side (matrix) of the membrane (Fig. 4B). The n-side cleft has a funnel shape 15 Å long (from aR176 to aD244 or aE162), 8 Å wide (from aD244 to aE162) and 16 Å deep (from aS165 to the C-terminus of subunit a) (Fig. 4B). It is bordered by polar and electrically charged residues of aH6 (aS240, aY241, aK243, aD244, aH249) and aH5 (aE162, aS165, aR169, aS172) (Fig. 4D), two of which (aE162 and aD244) are essential for moving protons out of the n-side cleft (25). The side chain of aR176, which is oriented towards the n-side cleft, helps through electrostatic interactions discharged subunit c monomer to rotate and be reloaded with a proton in the p-side cleft (25,43–45) (Fig. 4D). In addition to aL173, nine other residues are strictly conserved in subunits a: (i) aP106 allows the turn between aH3 and aH4; (ii) A183 ensures the shortest contact with the c-ring; (iii–vi) four hydrophilic residues line the bottom of the two half channels (aN100, aN180 and aQ230 in the entry one, and aY241 in the exit one); and (vii–ix) three leucine residues (aL173, aL177 and aL237) that together with three other highly but not strictly conserved hydrophobic residues (aI125, aV233 and aW234) form an hydrophobic plug that prevents proton shortage between the two channels (Fig. 4A and C).
Residue aL173 altered by the m.8993 T > G mutation is in aH5 just one α-helix turn from aR176 and aR169, and faces the essential glutamate in the c-ring involved in proton translocation, cE59 (Fig. 4D). Because of the proximity of aL173 to aR176 and cE59, it is not very surprising that its replacement with arginine compromises the function of FO (Fig. 4F). Without large conformational changes in subunit a, the aliphatic moiety of aR173 side chain preserves the hydrophobic barrier between the two proton half channels, whereas its guanidinium group increases electropositivity in the exit channel (Fig. 4D and F). As a result, proton discharge in the n-cleft and/or rotation of the c-ring might be expected to be affected. Consistent with this, the activity of the ATP synthase was substantially improved when aR173 was replaced with uncharged residues (aS173, aM173, aW173) (Fig. 4E). In apparent contradiction, the activity was also substantially recovered when aR173 was replaced with lysine (aK173). However, being further from the exit channel, the electric influence of the amine group of aK173, in particular on aD244 and aE162, is less important (Fig. 4G). Furthermore, the pKa of lysine is significantly lower than that of arginine, and within the local environment, the lysine may be uncharged. Partial recovery of ATP synthase activity by the aR169M/S changes further supports that the positive charge of aR173 compromises FO function (Fig. 4H). Notably, the aR169M mutation was also shown to compensate for another human pathogenic subunit a change in yeast (aW126R) presumably by preserving proper interactions between aR176 and cE59 (46). The proximal second-site reversions of aL173R (aA170P, aA170G) might similarly attenuate the detrimental effects of aR173 by structurally shifting its side chain.
Although substantially restored with aM173, aW173 or aK173, the measured ATP synthesis rate was diminished by 25–50% versus wild-type yeast (Table 2). For aM173 and aK173, the rates of ATP synthesis and electron transfer to oxygen were similarly diminished, indicating a reduction in the rate of proton transport through the FO possibly because of a reduced hydrophobicity around aR176 that makes the exit of protons towards the matrix and/or rotation of the c-ring less efficient. With aW173, there is a substantial (20–25%) drop in ATP production rate despite the preservation of a normal rate of electron transfer to oxygen. This indicates that the proton flow through the FO is only minimally affected. Indeed, as already mentioned above, yeast ATP synthase defective mutants show a slower rate of oxygen consumption only when FO activity is compromised (40). A possible explanation for the effects of aL173W is a partial proton shortage due to a distortion by the indole group of tryptophane of the hydrophobic barrier that seals the two half proton channels (Fig. 4E).
Conclusion
With the recently described cryo-EM structures of F1FO ATP synthases from various mitochondrial origins (25–27,44,45), it has become feasible to map at a molecular level discrete structural changes induced by mutations of this enzyme involved in human diseases. However, even with this information and a detailed knowledge of their biochemical consequences alone, it can be difficult to understand how ATP synthase function is compromised by the mutations. Genetic suppressors provide another input to help determine the mechanisms of pathogenic mutations. This approach is easily done in the yeast S. cerevisiae and this has proved successful in a number of previous studies (46–49). We herein applied it to the most common pathogenic mutation (m.8993 T > G, aL173R) of the mitochondrial ATP6 gene encoding the subunit a of ATP synthase. As we have shown (28), the leucine-to-arginine change (aL173R) leads to a 90% drop in yeast mitochondrial ATP synthesis, indicating a severe functional impairment of FO, but without defects in the assembly/stability of subunit a. Nine different amino acid changes of varying suppressor activity, at the original mutation site (aL173M, aL173S, aL173K and aL173W) or in another position of subunit a (aR169M, aR169S, aA170P, aA170G and aI216S) were here identified. The results support the hypothesis that the positively charged guanidine group of aR173 prevents the release of protons from the c-ring into the n-side hydrophilic cleft of FO likely due to unfavorable electrostatic interactions with two residues (aR176 and cE59), which are essential for the coupled movement of protons across the mitochondrial inner membrane with rotation of the c-ring.
Residue aL173 is not directly involved in proton conduction within the FO as it is located outside the n-side cleft and because its aliphatic side chain cannot exchange protons. Our results suggest, however, that aL173 is indirectly involved in the proton-conduction pathway. Indeed, none of the suppressors fully restored ATP synthase function. For aK173 and aM173, FO functions poorer compared to the wild-type enzyme, aW173 is responsible for a partial proton shortage between its two hydrophilic clefts, and respiratory growth is compromised with aS173. Using a similar suppressor genetics approach, we concluded that two other strictly conserved leucine residue of subunit a (aL237 and aL242) targeted by pathogenic mutations (m.9176 T > G, m.9176 T > C, and m.9191 T > C) also indirectly improve proton conduction within the wild-type FO (47,49).
In a previous study (48), we also provided evidence that another pathogenic mutation of subunit a (m.8969G > A; aS165N) leads to the establishment of hydrogen bonds that interferes with the proton conduction involving aE162. Remarkably, the m.8969G > A mutation was efficiently suppressed by long-range interactions upon mutations on the p-side of the membrane (at position 190 on aH5) (48). Similarly, the aL173R mutation was partially suppressed by a mutation (aI216S), which is close to residue 190 (on aH6) in the structure (Fig. 4A). This is an interesting finding that holds promise for the development of therapeutic molecules acting from the p-side of the membrane (thus without the need to cross the inner membrane).
Studies of human cells and tissues containing the m.8993 T > G mutation led to conflicting conclusions: a block in FO-mediated proton conduction (21), defects in ATP synthase assembly/stability (13,15), or impaired coupling between F1 and FO (18,50). The present yeast-based study provides evidence that the leucine-to-arginine change induced by this mutation prevents the release of protons from the c-ring due to unfavorable electrostatic interactions with two catalytic residues of FO (aR176 and cE59) without any defect in ATP synthase assembly/stability and proton shortage within the FO.
Materials and Methods
Growth media and genotypes
The media used for growing yeast were: YPGA (1% bacto yeast extract, 1% bacto peptone, 2% glucose, 60 mg/l adenine), YPGA10 (1% bacto yeast extract, 1% bacto peptone, 10% glucose, 60 mg/l adenine), YPGalA (1% bacto yeast extract, 1% bacto peptone, 2% galactose, 60 mg/l adenine), YPGlyA (1% bacto yeast extract, 1% bacto peptone, 2% glycerol, 60 mg/l adenine), WO (2% glucose, 0.67% yeast nitrogen base with ammonium sulfate from Difco), SP1: 0.1% glucose, 0.25% yeast extract, 50 mM potassium acetate. Solid media were made with 2% bacto agar (Difco, Becton Dickinson). The genotypes of the strains used in this study are listed in Table 4. Growth curves were determined with the Bioscreen C™ system (Piscataway, NJ).
Table 4.
Genotypes and origin of yeast strains
| Strain | Nuclear genotype | mtDNA | Source |
|---|---|---|---|
| MR6 | MAT𝑎 ade2–1 his3–11,15 trp1–1 leu2–3112 ura3–1 CAN1 arg8::HIS3 | ρ+ WT | (31) |
| MR14 | MAT𝑎 ade2–1 his3–11,15 trp1–1 leu2–3112 ura3–1 CAN1 arg8::HIS3 | ρ+atp6-L173R | (28) |
| DFS160 | MATα ade2–101 ura3–52 leu2∆ arg8::URA3 kar1–1 | ρ0 | (64) |
| KL14-4A | MAT𝑎 his1 trp2 | ρ+oli1r-1-1 (atp9-L53F) capr-1-321 parr-1-454 | (65) This study |
| SDC31 | MATα ade2–101 ura3–52 leu2∆ arg8::URA3 kar1–1 | ρ−atp6-L173R | (28) |
| MR11 | MATα/𝑎 ade2–101/+ ura3–52/+ leu2∆/+ arg8::URA3/+ kar1–1/+ +/his1 +/trp2 | ρ+oli1r-1-1 (atp9-L53F) capr-1-321 parr-1-454 | This study |
| MR12 | MATα/𝑎 ade2–101/+ ura3–52/+ leu2∆/+ arg8::URA3/+ kar1–1/+ +/his1 +/trp2 | ρ+atp6-L173R oli1r-1-1 (atp9-L53F) capr-1-321 parr-1-454 |
This study |
| RMR12/1 | Revertant isolated from MR12 | ρ+atp6-L173M | This study |
| RMR12/6 | Revertant isolated from MR12 | ρ+atp6-L173K | This study |
| RMR12/11 | Revertant isolated from MR12 | ρ+atp6-L173S | This study |
| RMR12/8 | Revertant isolated from MR12 | ρ+atp6-L173R + R169S | This study |
| RMR12/13 | Revertant isolated from MR12 | ρ+atp6-L173R + A170P | This study |
| RMR12/20 | Revertant isolated from MR12 | ρ+atp6-L173R + A170G | This study |
| RMR12/5 | Revertant isolated from MR12 | ρ+atp6-L173R + I216S | This study |
| RMR14/4 | Revertant isolated from MR14 | ρ+atp6-L173W | This study |
| RMR14/6 | Revertant isolated from MR14 | ρ+atp6-L173K | This study |
| RMR14/12 | Revertant isolated from MR14 | ρ+atp6-L173M | This study |
| RMR14/48 | Revertant isolated from MR14 | ρ+atp6-L173R + R169M | This study |
Construction of the aL173R diploid strain (MR12)
As described in (28), the plasmid (pSDC31) harboring the ATP6 gene with the aL173R mutation (TTA183AGA) was introduced into the mitochondria of the strain DFS160 that is entirely devoid of mtDNA (ρ0). The resulting ρ− synthetic strain (SDC31) was crossed to the ρ+ strain KL14-4A. Diploids cells were selected in a minimal glucose medium (WO) and allowed to divide for at least 15 generations for homoplasmy of ρ + mtDNA recombinants carrying the atp6-L173R mutation. Single diploid clones were selected on solid minimal medium (WO) and tested for growth on rich glycerol medium (YPGlyA). Out of 1000 clones, 5 (called MR12) carried the aL173R mutation and all 5 grew very slowly on glycerol in comparison with the control diploid strain (MR11) obtained by crossing DFS160 to KL14-4A.
Selection of genetic suppressors of aL173R
Genetic suppressors of aL173R were selected from diploid strain MR12 and a haploid strain (MR14) carrying the same mutation that was previously described (28). To ensure genetic independence of the isolates, the strains were streaked for single colonies on YPGA plates and dozens of subclones were picked up and individually grown for two days in 10% glucose (YPGA). Glucose was removed from the cultures by washings twice with water and 108 cells from each culture were spread on rich glycerol (YPGlyA) plates. The plates were incubated at 28°C for at least fifteen days. A maximum two revertants per plate were selected for further analysis. The revertants were purified by subcloning on YPGA and their ATP6 gene was PCR-amplified and sequenced with two pairs of primers: (i) 5’-TAATATACGGGGGTGGGTCCCTCAC (Forward, from nucleotide position −100 upstream of the ATP6 start codon) and 5’-CTGCATCTTTTAAATATGATGCTG (Reverse, from nucleotide position +743 downstream of the ATP6 start codon); (ii) 5’-GTATGATTCCATACTCATTTG (Forward, from nucleotide position +337 downstream of the ATP6 start codon) and 5’-GGGCCGAACTCCGAAGGAGTAAG (reverse, from nucleotide position +218 downstream of the ATP6 stop codon).
Miscellaneous procedures
Mitochondria were isolated by the previously described enzymatic method (51) from yeast cells grown until middle exponential phase (2–3 × 107 cells/mL) in rich galactose medium (YPGalA) at 28°C. Protein concentration was determined by the Lowry method (52) in the presence of 5% SDS. Oxygen consumption rates were measured with a Clark electrode using 150 μg/mL of mitochondrial proteins in a respiration buffer (0.65 M mannitol, 0.36 mM ethylene glycol tetra-acetic acid, 5 mM Tris/phosphate, 10 mM Tris/maleate pH 6.8) as in (53). The additions were 4 mM NADH, 12.5 mM ascorbate (Asc), 1.4 mM N, N, N′, N′-tetramethyl-p-phenylenediamine (TMPD), 150 μM ADP and 4 μM carbonyl cyanide m-chlorophenylhydrazone (CCCP). The transmembrane potential (ΔΨ) was evaluated in respiration buffer and 150 μg/mL of mitochondrial protein, by monitoring fluorescence quenching of Rhodamine 123 (25 μg/mL) using an SAFAS Monaco spectrophotometer as described in (54). The additions were 10 μL ethanol (EtOH), 75 μM ADP, 2 mM potassium cyanide (KCN), 0.2 mM ATP, 4 μg/mL oligomycin (oligo) and 4 μM CCCP. ATP synthesis rate measurements were performed with 150 μg/mL of mitochondrial proteins in respiration buffer supplemented with 4 mM NADH and 1 mM ADP in a thermostatically controlled chamber at 28°C, as described in (31). Aliquots of 50 μL were withdrawn from the oxygraphy cuvette every 15 s and the reaction was stopped with 7% (w/v) perchloric acid, 25 mM EDTA. The samples were then neutralized to pH 6.8 by adding 2 M KOH/0.3 M MOPS, and ATP was quantified using a luciferin/luciferase assay (ATPlite kit from Perkin Elmer) and the CLARIO star microplate reader (from BMG LABTECH). The participation of the F1FO-ATP synthase in ATP production was assessed by adding oligomycin (20 μg/mg of protein). ATPase activity measurements were performed with 150 μg/mL of mitochondrial proteins in Somlo ATPase buffer (0.2 M KCl, 3 mM MgCl2, 10 mM Tris–HCl, adjusted to pH 8.4) as described (55). The reaction was started by the addition of 50 μL 0.1 M ATP. After 2 min, the reaction was stopped and proteins were precipitated by adding TCA 100% and ATP hydrolysis was measured by measuring inorganic phosphate using ammonium molybdate solution and measuring the absorbance at 610 nm (56). Oligomycin sensitive ATP hydrolysis of F1FO-ATPase was assessed by adding oligomycin (2 μg/mL). Blue native polyacrylamide gel electrophoresis (BN-PAGE) analyses was done as described (57). Briefly, mitochondrial extracts solubilized with 2 g digitonin/g protein were separated in a 3–12% acrylamide continuous gradient gel. ATP synthase complexes were detected in-gel by their ATPase activity as described (58), or by western blot after transferred to polyvinylidene difluoride (PVDF) using polyclonal antibodies against Atp6/subunit a (a gift from J. Velours), Atp1/subunit α (a gift from J. Velours) and Atp9/subunit c after 1:5000 dilution (59). SDS-PAGE analyses of mitochondrial proteins was performed according to (60). 1:500 dilutions of monoclonal antibodies against Cox2 and Porin (Molecular Probes) were used. The tested proteins were revealed using peroxidase-labeled antibodies (Promega) at a 1:5000 dilution and the ECL reagent of Pierce™ ECL western blotting substrate (ThermoScientific). Immunological signal quantification was performed using ImageJ (61).
Amino-acid alignments and structural modeling
Multiple sequence alignment of ATP synthase a-subunits of various origins was performed using Clustal Omega (62). Molecular views of subunit a and c10-ring were obtained from the dimeric Fo domain of S. cerevisiae ATP synthase (pdb_id: 6b2z; 3.6 Å resolution (25)). All structure figures were drawn using PyMOL molecular graphic system (63).
Supplementary Material
Contributor Information
Xin Su, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
Alain Dautant, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
Malgorzata Rak, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
François Godard, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
Nahia Ezkurdia, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
Marine Bouhier, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
Maïlis Bietenhader, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
David M Mueller, Center for Genetic Diseases, Chicago Medical School, Rosalind Franklin University, 3333 Green Bay Rd, North Chicago, IL, 60064, USA.
Roza Kucharczyk, Institute of Biochemistry and Biophysics, Polish Academy of Sciences, 00090 Warsaw, Poland.
Jean-Paul di Rago, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
Déborah Tribouillard-Tanvier, University Bordeaux, CNRS, IBGC, UMR 5095, F-33000 Bordeaux, France.
Funding
A Ph.D. fellowship from China Scholarship Council (CSC) to X.S.; PhD fellowships from CNRS and Retina foundation, respectively to M.Bi. and M.R.; the Association Française contre les Myopathies (#22382 to D.T.T.) and J.P.dR.; the National Science Center of Poland (NSCP 2016/23/B/NZ3/02098 to R.K.); from National Institutes of Health (R35GM131731 to D.M.M).
Author contributions
X.S., M.R., M.Bi., R.K. and N.E.G. isolated, sequenced and biochemically characterized yeast strains under the supervision of D.T.T. and J.P.dR. and with the technical support of M.Bo. and F.G. a.d. performed the structural modeling analyses. X.S., a.d., R.K., M.Bi., D.T.T., J.P.dR. and D.M.M. analyzed the data. X.S., a.d., D.T.T., J.P.dR. and D.M.M. wrote the paper. D.T.T. and J.P.dR. designed the research. Conflict of interest statement: The authors declare no competing or financial interests.
References
- 1. Zeviani, M. and Di Donato, S. (2004) Mitochondrial disorders. Brain, 127, 2153–2172. [DOI] [PubMed] [Google Scholar]
- 2. Thompson, K., Collier, J.J., Glasgow, R.I.C., Robertson, F.M., Pyle, A., Blakely, E.L., Alston, C.L., Olahova, M., McFarland, R. and Taylor, R.W. (2020) Recent advances in understanding the molecular genetic basis of mitochondrial disease. J. Inherit. Metab. Dis., 43, 36–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Saraste, M. (1999) Oxidative phosphorylation at the fin de siècle. Science, 283, 1488–1493. [DOI] [PubMed] [Google Scholar]
- 4. Boyer, P.D. (1997) The ATP synthase—a splendid molecular machine. Annu. Rev. Biochem., 66, 717–749. [DOI] [PubMed] [Google Scholar]
- 5. Richter, C. (1992) Reactive oxygen and DNA damage in mitochondria. Mutation Research/DNAging, 275, 249–255. [DOI] [PubMed] [Google Scholar]
- 6. Croteau, D.L., Stierum, R.H. and Bohr, V.A. (1999) Mitochondrial DNA repair pathways. Mutation Research/DNA Repair, 434, 137–148. [DOI] [PubMed] [Google Scholar]
- 7. Moraes, C.T. (2001) What regulates mitochondrial DNA copy number in animal cells? Trends Genet., 17, 199–205. [DOI] [PubMed] [Google Scholar]
- 8. D’Souza, A.D., Parikh, N., Kaech, S.M. and Shadel, G.S. (2007) Convergence of multiple signaling pathways is required to coordinately up-regulate mtDNA and mitochondrial biogenesis during T cell activation. Mitochondrion, 7, 374–385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Bonnefoy, N. and Fox, T.D. (2001) Genetic transformation of Saccharomyces cerevisiae mitochondria. Methods Cell Biol., 65, 381–396. [DOI] [PubMed] [Google Scholar]
- 10. Okamoto, K., Perlman, P.S. and Butow, R.A. (1998) The sorting of mitochondrial DNA and mitochondrial proteins in zygotes: preferential transmission of mitochondrial DNA to the medial bud. J. Cell Biol., 142, 613–623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Tatuch, Y., Christodoulou, J., Feigenbaum, A., Clarke, J.T., Wherret, J., Smith, C., Rudd, N., Petrova-Benedict, R. and Robinson, B.H. (1992) Heteroplasmic mtDNA mutation (T----G) at 8993 can cause Leigh disease when the percentage of abnormal mtDNA is high. Am. J. Hum. Genet., 50, 852–858. [PMC free article] [PubMed] [Google Scholar]
- 12. Trounce, I., Neill, S. and Wallace, D.C. (1994) Cytoplasmic transfer of the mtDNA nt 8993 T-->G (ATP6) point mutation associated with Leigh syndrome into mtDNA-less cells demonstrates cosegregation with a decrease in state III respiration and ADP/O ratio. Proc. Natl. Acad. Sci. U. S. A., 91, 8334–8338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Houstek, J., Klement, P., Hermanska, J., Houstkova, H., Hansikova, H., Van den Bogert, C. and Zeman, J. (1995) Altered properties of mitochondrial ATP-synthase in patients with a T-->G mutation in the ATPase 6 (subunit a) gene at position 8993 of mtDNA. Biochim. Biophys. Acta, 1271, 349–357. [DOI] [PubMed] [Google Scholar]
- 14. Garcia, J.J., Ogilvie, I., Robinson, B.H. and Capaldi, R.A. (2000) Structure, functioning, and assembly of the ATP synthase in cells from patients with the T8993G mitochondrial DNA mutation. Comparison with the enzyme in rho(0) cells completely lacking mtdna. J. Biol. Chem., 275, 11075–11081. [DOI] [PubMed] [Google Scholar]
- 15. Nijtmans, L.G., Henderson, N.S., Attardi, G. and Holt, I.J. (2001) Impaired ATP synthase assembly associated with a mutation in the human ATP synthase subunit 6 gene. J. Biol. Chem., 276, 6755–6762. [DOI] [PubMed] [Google Scholar]
- 16. Mattiazzi, M., Vijayvergiya, C., Gajewski, C.D., DeVivo, D.C., Lenaz, G., Wiedmann, M. and Manfredi, G. (2004) The mtDNA T8993G (NARP) mutation results in an impairment of oxidative phosphorylation that can be improved by antioxidants. Hum. Mol. Genet., 13, 869–879. [DOI] [PubMed] [Google Scholar]
- 17. Carrozzo, R., Rizza, T., Stringaro, A., Pierini, R., Mormone, E., Santorelli, F.M., Malorni, W. and Matarrese, P. (2004) Maternally-inherited Leigh syndrome-related mutations bolster mitochondrial-mediated apoptosis. J. Neurochem., 90, 490–501. [DOI] [PubMed] [Google Scholar]
- 18. Sgarbi, G., Baracca, A., Lenaz, G., Valentino, L.M., Carelli, V. and Solaini, G. (2006) Inefficient coupling between proton transport and ATP synthesis may be the pathogenic mechanism for NARP and Leigh syndrome resulting from the T8993G mutation in mtDNA. Biochem. J., 395, 493–500. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Baracca, A., Sgarbi, G., Mattiazzi, M., Casalena, G., Pagnotta, E., Valentino, M.L., Moggio, M., Lenaz, G., Carelli, V. and Solaini, G. (2007) Biochemical phenotypes associated with the mitochondrial ATP6 gene mutations at nt8993. Biochim. Biophys. Acta, 1767, 913–919. [DOI] [PubMed] [Google Scholar]
- 20. Morava, E., Rodenburg, R.J., Hol, F., deVries, M., Janssen, A., van denHeuvel, L., Nijtmans, L. and Smeitink, J. (2006) Clinical and biochemical characteristics in patients with a high mutant load of the mitochondrial T8993G/C mutations. Am. J. Med. Genet. A, 140, 863–868. [DOI] [PubMed] [Google Scholar]
- 21. Cortes-Hernandez, P., Vazquez-Memije, M.E. and Garcia, J.J. (2007) ATP6 homoplasmic mutations inhibit and destabilize the human F1F0-ATP synthase without preventing enzyme assembly and oligomerization. J. Biol. Chem., 282, 1051–1058. [DOI] [PubMed] [Google Scholar]
- 22. Pallotti, F., Baracca, A., Hernandez-Rosa, E., Walker, W.F., Solaini, G., Lenaz, G., Melzi D'Eril, G.V., Dimauro, S., Schon, E.A. and Davidson, M.M. (2004) Biochemical analysis of respiratory function in cybrid cell lines harbouring mitochondrial DNA mutations. Biochem. J., 384, 287–293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Vazquez-Memije, M.E., Shanske, S., Santorelli, F.M., Kranz-Eble, P., DeVivo, D.C. and DiMauro, S. (1998) Comparative biochemical studies of ATPases in cells from patients with the T8993G or T8993C mitochondrial DNA mutations. J. Inherit. Metab. Dis., 21, 829–836. [DOI] [PubMed] [Google Scholar]
- 24. Zhou, A., Rohou, A., Schep, D.G., Bason, J.V., Montgomery, M.G., Walker, J.E., Grigorieff, N. and Rubinstein, J.L. (2015) Structure and conformational states of the bovine mitochondrial ATP synthase by cryo-EM. elife, 4, e10180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Guo, H., Bueler, S.A. and Rubinstein, J.L. (2017) Atomic model for the dimeric FO region of mitochondrial ATP synthase. Science, 358, 936–940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Srivastava, A.P., Luo, M., Zhou, W., Symersky, J., Bai, D., Chambers, M.G., Faraldo-Gomez, J.D., Liao, M. and Mueller, D.M. (2018) High-resolution cryo-EM analysis of the yeast ATP synthase in a lipid membrane. Science, 360. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Kühlbrandt, W. (2019) Structure and mechanisms of F-type ATP synthases. Annu. Rev. Biochem., 88, 515–549. [DOI] [PubMed] [Google Scholar]
- 28. Rak, M., Tetaud, E., Duvezin-Caubet, S., Ezkurdia, N., Bietenhader, M., Rytka, J. and diRago, J.P. (2007) A yeast model of the neurogenic ataxia retinitis pigmentosa (NARP) T8993G mutation in the mitochondrial ATP synthase-6 gene. J. Biol. Chem., 282, 34039–34047. [DOI] [PubMed] [Google Scholar]
- 29. Dautant, A., Meier, T., Hahn, A., Tribouillard-Tanvier, D., diRago, J.P. and Kucharczyk, R. (2018) ATP synthase diseases of mitochondrial genetic origin. Front. Physiol., 9, 329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Michon, T., Galante, M. and Velours, J. (1988) NH2-terminal sequence of the isolated yeast ATP synthase subunit 6 reveals post-translational cleavage. Eur. J. Biochem., 172, 621–625. [DOI] [PubMed] [Google Scholar]
- 31. Rak, M., Tetaud, E., Godard, F., Sagot, I., Salin, B., Duvezin-Caubet, S., Slonimski, P.P., Rytka, J. and diRago, J.P. (2007) Yeast cells lacking the mitochondrial gene encoding the ATP synthase subunit 6 exhibit a selective loss of complex IV and unusual mitochondrial morphology. J. Biol. Chem., 282, 10853–10864. [DOI] [PubMed] [Google Scholar]
- 32. Tzagoloff, A., Barrientos, A., Neupert, W. and Herrmann, J.M. (2004) Atp10p assists assembly of Atp6p into the F0 unit of the yeast mitochondrial ATPase. J. Biol. Chem., 279, 19775–19780. [DOI] [PubMed] [Google Scholar]
- 33. Godard, F., Tetaud, E., Duvezin-Caubet, S. and diRago, J.P. (2011) A genetic screen targeted on the FO component of mitochondrial ATP synthase in Saccharomyces cerevisiae. J. Biol. Chem., 286, 18181–18189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Rak, M., Zeng, X., Briere, J.J. and Tzagoloff, A. (2009) Assembly of F0 in Saccharomyces cerevisiae. Biochim. Biophys. Acta, 1793, 108–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Sanchirico, M.E., Fox, T.D. and Mason, T.L. (1998) Accumulation of mitochondrially synthesized Saccharomyces cerevisiae Cox2p and Cox3p depends on targeting information in untranslated portions of their mRNAs. EMBO J., 17, 5796–5804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Kucharczyk, R., Ezkurdia, N., Couplan, E., Procaccio, V., Ackerman, S.H., Blondel, M. and diRago, J.P. (2010) Consequences of the pathogenic T9176C mutation of human mitochondrial DNA on yeast mitochondrial ATP synthase. Biochim. Biophys. Acta, 1797, 1105–1112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Kucharczyk, R., Salin, B. and diRago, J.P. (2009) Introducing the human Leigh syndrome mutation T9176G into Saccharomyces cerevisiae mitochondrial DNA leads to severe defects in the incorporation of Atp6p into the ATP synthase and in the mitochondrial morphology. Hum. Mol. Genet., 18, 2889–2898. [DOI] [PubMed] [Google Scholar]
- 38. Kucharczyk, R., Rak, M. and diRago, J.-P. (2009) Biochemical consequences in yeast of the human mitochondrial DNA 8993T>C mutation in the ATPase6 gene found in NARP/MILS patients. BBA - Mol. Cell Res., 1793, 817–824. [DOI] [PubMed] [Google Scholar]
- 39. Duvezin-Caubet, S., Caron, M., Giraud, M.F., Velours, J. and diRago, J.P. (2003) The two rotor components of yeast mitochondrial ATP synthase are mechanically coupled by subunit delta. Proc. Natl. Acad. Sci. U. S. A., 100, 13235–13240. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Su, X., Rak, M., Tetaud, E., Godard, F., Sardin, E., Bouhier, M., Gombeau, K., Caetano-Anollés, D., Salin, B., Chen, H.et al. (2019) Deregulating mitochondrial metabolite and ion transport has beneficial effects in yeast and human cellular models for NARP syndrome. Hum. Mol. Genet., in press. [DOI] [PubMed] [Google Scholar]
- 41. Kucharczyk, R., Giraud, M.F., Brethes, D., Wysocka-Kapcinska, M., Ezkurdia, N., Salin, B., Velours, J., Camougrand, N., Haraux, F. and diRago, J.P. (2013) Defining the pathogenesis of human mtDNA mutations using a yeast model: the case of T8851C. Int. J. Biochem. Cell Biol., 45, 130–140. [DOI] [PubMed] [Google Scholar]
- 42. Venard Rene’e, D.B.t., Giraud Marie-France, Vaillier Jacques, Velours Jean, and Haraux Francis. (2003) Investigation of the role and mechanism of IF1 and STF1 proteins, twin inhibitory peptides which interact with the yeast mitochondrial ATP synthase. Biochemistry, 42,7626.–. [DOI] [PubMed] [Google Scholar]
- 43. Klusch, N., Murphy, B.J., Mills, D.J., Yildiz, O. and Kuhlbrandt, W. (2017) Structural basis of proton translocation and force generation in mitochondrial ATP synthase. elife, 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Hahn, A., Vonck, J., Mills, D.J., Meier, T. and Kuhlbrandt, W. (2018) Structure, mechanism, and regulation of the chloroplast ATP synthase. Science, 360. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Murphy, B.J., Klusch, N., Langer, J., Mills, D.J., Yildiz, O. and Kuhlbrandt, W. (2019) Rotary substates of mitochondrial ATP synthase reveal the basis of flexible F1-Fo coupling. Science, 364. [DOI] [PubMed] [Google Scholar]
- 46. Kucharczyk, R., Dautant, A., Gombeau, K., Godard, F., Tribouillard-Tanvier, D. and diRago, J.P. (2019) The pathogenic MT-ATP6 m.8851T>C mutation prevents proton movements within the n-side hydrophilic cleft of the membrane domain of ATP synthase. Biochim. Biophys. Acta Bioenerg., 1860, 562–572. [DOI] [PubMed] [Google Scholar]
- 47. Kucharczyk, R., Dautant, A., Godard, F., Tribouillard-Tanvier, D. and diRago, J.P. (2019) Functional investigation of an universally conserved leucine residue in subunit a of ATP synthase targeted by the pathogenic m.9176T>G mutation. Biochim. Biophys. Acta Bioenerg., 1860, 52–59. [DOI] [PubMed] [Google Scholar]
- 48. Skoczen, N., Dautant, A., Binko, K., Godard, F., Bouhier, M., Su, X., Lasserre, J.P., Giraud, M.F., Tribouillard-Tanvier, D., Chen, H.et al. (2018) Molecular basis of diseases caused by the mtDNA mutation m.8969G>a in the subunit a of ATP synthase. Biochim. Biophys. Acta Bioenerg., 1859, 602–611. [DOI] [PubMed] [Google Scholar]
- 49. Su, X., Dautant, A., Godard, F., Bouhier, M., Zoladek, T., Kucharczyk, R., diRago, J.P. and Tribouillard-Tanvier, D. (2020) Molecular basis of the pathogenic mechanism induced by the m.9191T>C mutation in mitochondrial ATP6 gene. Int. J. Mol. Sci., 21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Baracca, A., Barogi, S., Carelli, V., Lenaz, G. and Solaini, G. (2000) Catalytic activities of mitochondrial ATP synthase in patients with mitochondrial DNA T8993G mutation in the ATPase 6 gene encoding subunit a. J. Biol. Chem., 275, 4177–4182. [DOI] [PubMed] [Google Scholar]
- 51. Guérin, B., Labbe, P. and Somlo, M. (1979) Methods in Enzymology. Elsevier, Vol. 55, pp. 149–159. [DOI] [PubMed] [Google Scholar]
- 52. Lowry, O.H., Rosebrough, N.J., Farr, A.L. and Randall, R.J. (1951) Protein measurement with the Folin phenol reagent. J. Biol. Chem., 193, 265–275. [PubMed] [Google Scholar]
- 53. Rigoulet, M. and Guerin, B. (1979) Phosphate transport and ATP synthesis in yeast mitochondria: effect of a new inhibitor: the tribenzylphosphate. FEBS Lett., 102, 18–22. [DOI] [PubMed] [Google Scholar]
- 54. Emaus, R.K., Grunwald, R. and Lemasters, J.J. (1986) Rhodamine 123 as a probe of transmembrane potential in isolated rat-liver mitochondria: spectral and metabolic properties. Biochim. Biophys. Acta, 850, 436–448. [DOI] [PubMed] [Google Scholar]
- 55. Somlo, M. (1968) Induction and repression of mitochondrial ATPase in yeast. Eur. J. Biochem., 5, 276–284. [DOI] [PubMed] [Google Scholar]
- 56. Lanzetta, P.A., Alvarez, L.J., Reinach, P.S. and Candia, O.A. (1979) An improved assay for nanomole amounts of inorganic phosphate. Anal. Biochem., 100, 95–97. [DOI] [PubMed] [Google Scholar]
- 57. Schagger, H. and vonJagow, G. (1991) Blue native electrophoresis for isolation of membrane protein complexes in enzymatically active form. Anal. Biochem., 199, 223–231. [DOI] [PubMed] [Google Scholar]
- 58. Grandier-Vazeille, X. and Guerin, M. (1996) Separation by blue native and colorless native polyacrylamide gel electrophoresis of the oxidative phosphorylation complexes of yeast mitochondria solubilized by different detergents: specific staining of the different complexes. Anal. Biochem., 242, 248–254. [DOI] [PubMed] [Google Scholar]
- 59. Arselin, G., Vaillier, J., Graves, P.V. and Velours, J. (1996) ATP synthase of yeast mitochondria. Isolation of the subunit h and disruption of the ATP14 gene. J. Biol. Chem., 271, 20284–20290. [DOI] [PubMed] [Google Scholar]
- 60. Laemmli, U.K. (1970) Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature, 227, 680–685. [DOI] [PubMed] [Google Scholar]
- 61. Gassmann, M., Grenacher, B., Rohde, B. and Vogel, J. (2009) Quantifying western blots: pitfalls of densitometry. Electrophoresis, 30, 1845–1855. [DOI] [PubMed] [Google Scholar]
- 62. Sievers, F., Wilm, A., Dineen, D., Gibson, T.J., Karplus, K., Li, W., Lopez, R., McWilliam, H., Remmert, M., Soding, J.et al. (2011) Fast, scalable generation of high-quality protein multiple sequence alignments using clustal omega. Mol. Syst. Biol., 7, 539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Schrödinger, LLC. (2002) The PyMOL Molecular Graphics System, Version 0.99. DeLano Scientific, San Carlos, CA, USA.
- 64. Steele, D.F., Butler, C.A. and Fox, T.D. (1996) Expression of a recoded nuclear gene inserted into yeast mitochondrial DNA is limited by mRNA-specific translational activation. Proc. Natl. Acad. Sci., 93, 5253–5257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Colson, A.M. and Slonimski, P.P. (1979) Genetic localization of diuron- and mucidin-resistant mutants relative to a group of loci of the mitochondrial DNA controlling coenzyme QH2-cytochrome c reductase in Saccharomyces cerevisiae. Mol. Gen. Genet., 167, 287–298. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
