Summary
Following the Goldilocks principle, mitochondria size must be ‘just right’. Mitochondria balance division and fusion to avoid becoming too big or too small. Defects in this balance produce dysfunctional mitochondria in human diseases. Mitochondrial safeguard (MitoSafe) is a defense mechanism that protects mitochondria against extreme enlarging by suppressing fusion in mammalian cells. In MitoSafe, hyperfused mitochondria elicit flickering – short pulses of mitochondrial depolarization. Flickering activates an inner membrane protease, Oma1, which in turn proteolytically inactivates a mitochondrial fusion protein, Opa1. The mechanisms underlying flickering are unknown. Using a live-imaging screen, we identified Slc25a3 (a mitochondrial carrier transporting phosphate and copper) as necessary for flickering and Opa1 cleavage. Remarkably, copper, but not phosphate, is critical for flickering. Furthermore, we found that two copper-containing mitochondrial enzymes, superoxide dismutase 1 and cytochrome c oxidase, regulate flickering. Our data identify an unforeseen mechanism linking copper, redox homeostasis, and membrane flickering in mitochondrial defense against deleterious fusion.
Keywords: Mitochondria, Fusion, Division, Stress response, Transporter
Graphical Abstract

eTOC
Murata et al. report that mitochondrial copper, transported by Slc25a3, regulates the balance between mitochondrial fusion and division by suppressing fusion when mitochondria become hyper-fused. This function of copper is mediated by two copper-containing enzymes: superoxide dismutase 1 and cytochrome c oxidase.
Introduction
Mitochondria grow by importing proteins and lipids and control their morphology by dividing and fusing their membranes 1,2. A balance between division and fusion determines the mitochondrial size, number, and connectivity in cells 3–6. Excess fusion results in the elongation and enlargement of mitochondria, which become defective in intracellular transport or mitophagy, while excess division produces extremely small mitochondria, many of which lack mtDNA. Structural abnormalities are associated with mitochondrial energy deficiency, aberrant distribution, and impaired autophagic turnover in many human diseases, including brain development defects, muscle atrophy, cardiovascular diseases, cancers, and metabolic syndrome 7–11. Therefore, like in the story of Goldilocks and the Three Bears, mitochondria need to be ‘just right’, not too small and not too big.
Evolutionarily conserved dynamin-related GTPases mediate mitochondrial division and fusion. Mitochondria are severed by dynamin-related protein 1 (Drp1), while mitochondria are fused by mitofusin 1 and 2 (Mfn1 and 2) at the outer membrane and optic atrophy 1 (Opa1) at the inner membrane 1,2,12–18. The activities of these GTPases must be coordinated to prevent unbalanced mitochondrial division and fusion. We recently identified a defense mechanism, termed mitochondrial safeguard (MitoSafe). MitoSafe regulates mitochondrial fusion when mitochondria are hyper-fused either due to the genetic loss of Drp1 or its receptor, or due to metabolic stress caused by depleting methionine and choline in mice 19–22. MitoSafe prevents deleterious, extreme fusion that impairs mitochondrial respiratory function, at both the outer membrane and inner membranes 19. At the outer membrane, a ubiquitin E3 ligase (Parkin) and a mitochondrial kinase that phosphorylates parkin and ubiquitin (PINK1) promote proteasomal degradation of Mfn1 and Mfn2 20,21. At the inner membrane, a metalloprotease (Oma1), cleaves Opa1 and suppresses inner membrane fusion 19.
In MitoSafe, Oma1 is activated by repeated, transient decreases of the inner membrane potential, termed flickering, without losing the overall membrane potential (Fig. 1A) 19. Flickering has been reported in several cell types, including smooth muscle cells 23, cardiomyocytes 24, and a neuroblastoma cell line 25; however, its function remained largely unknown. The flickering occurs, independently of mitochondrial respiration, calcium signaling, or a mitochondrial permeability transition pore 19,26,27. Remarkably, in hyper-fused mitochondria, the frequency of flickering is greatly accelerated 19,26,27. Thus, MitoSafe represents the critical biological process that could be considered a function of flickering 19. However, it is unknown how elongated mitochondria induce flickering.
Figure 1. A knockdown screen identifies that Slc25a3 is critical for flickering and Opa1 cleavage in Drp1-KO cells.

(A) Drp1-KO MEFs expressing matrix-targeted Su9-GFP were viewed by laser scanning confocal microscopy for 30 min with 10-s intervals in the presence of TMRE. The arrows indicate mitochondria that showed flickering. Three frames from the time-lapse analysis are shown. Scale bar, 10 μm. (B) A list of genes tested. (C) 30 genes were individually knocked down in Drp1-KO cells using esiRNA and analyzed for flickering by TMRE staining and live-cell imaging. Knockdown of four genes highlighted in yellow in (B) decreased the frequency of flickering. (D) Drp1-KO MEFs were transduced by lentiviruses carrying shRNAs for each of the four candidates for 1-2 weeks. Cells were stained with TMRE and viewed by laser scanning confocal microscopy for 30 min with 10-s intervals. The percentage of cells that showed flickering are presented (average ± SD, n = 3). (E) Slc25a3 was knocked out using CRISPR in Drp1-KO MEFs. Flickering frequency is quantified (average ± SD, n = 3). (F) The mitochondrial membrane potential was measured in Drp1-KO and Drp1Slc25a3-KO MEFs using MitoLite and flow cytometry. As a negative control, Drp1-KO MEFs were treated with 10 μM FCCP. Significance was calculated using ANOVA with post-hoc Tukey in (D) and Student’s t-test in (E): **p<0.01. See also Figures S1 and S2.
Results
A live-cell imaging-based RNA interference screen identifies Slc25a3 as critical for flickering in Drp1-KO cells
To uncover the mechanism that promotes flickering upon mitochondrial hyperfusion in Drp1-KO cells, we selected mitochondrial proteins based on the mitochondrial proteome (MitoCarta3.0) and further narrowed our choice to known or potential carrier proteins that may transport small molecules or ions 28. We also included some membrane proteins with unknown functions 28. We individually knocked down 30 selected genes in Drp1-KO mouse embryonic fibroblasts (MEFs) or Drp1-KO HEK293T cells in the following two screens (Fig. 1B). In the first screening, we transfected Drp1-KO cells with a mixture of endoribonuclease-prepared siRNAs (esiRNAs), which is designed to target a single gene. This approach ensures specific and effective gene knockdown 29. Five days post-transfection, the cells were stained with the membrane potential-dependent dye, tetramethylrhodamine ethyl ester (TMRE), and subjected to live-cell imaging for 30 minutes. The membrane potential was not globally decreased by the knockdown of the 30 genes (Fig. S1). We scored cells that exhibited flickering during the 30-minute observation period. Typically, approximately 60-70% of Drp1-KO cells exhibited flickering (Fig. 1A and C). This knockdown experiment was conducted twice, with a threshold set at an approximate 50% reduction in flickering frequency compared to Drp1-KO cells treated with scramble esiRNAs. Four genes (Ccdc51, Mpv17, Slc25a3, and Slc25a40) met this criterion in both trials (Fig. 1C). In the second screen, we knocked down each candidate gene in Drp1-KO MEFs using shRNAs and monitored flickering in three separate experiments (Fig. 1D). The knockdown of Slc25a3 was found to significantly decrease the frequency of flickering in Drp1-KO MEFs (Fig. 1D).
Slc25a3 is a member of the mitochondrial solute carrier family and located in the inner membrane. In humans, mutations in Slc25a3 causes fatal mitochondrial defects with lactic acidosis, hypertrophic cardiomyopathy, and neonatal hypotonia 30,31. Slc25a3 was initially identified as a phosphate carrier, and recent studies have shown that Slc25a3 also transports copper 32,33. To further confirm the role of Slc25a3 in flickering, we knocked out Slc25a3 in Drp1-KO MEFs using CRISPR gene editing (Fig. S2). Consistent with the knockdown data, knockout of Slc25a3 significantly reduced the flickering frequency in Drp1-KO MFEs (Fig. 1E). This is not due to a drop in the membrane potential since Drp1Slc25a3-KO cells maintain a membrane potential similar to Drp1-KO MEFs (Fig. 1F).
Slc25a3 is important for enhanced OPA1 cleavage in Drp1-KO cells
We have previously shown that flickering promotes the proteolytic processing of Opa1 by the metalloprotease Oma119. The Opa1 long isoforms, L1 and L2, are cleaved to form the short forms, S3 and S5, respectively 2,19,34–36. This Oma1-mediated Opa1 cleavage is enhanced in Drp1-KO cells 19 (Fig. 2A and B). To determine whether the loss of Slc25a3 affects Opa1 processing, we analyzed Opa1 in WT, Drp1-KO, Slc25a3-KO, and Drp1Slc25a3-KO MEFs using Western blotting. We found that levels of L1 and L2 were significantly increased in Drp1Slc25a3-KO MEFs compared to Drp1-KO MEFs (Fig. 2A and B). In contrast to Opa1, decreased levels of Mfn1 and Mfn2 in Drp1-KO MEFs were not restored in Drp1Slc25a3-KO MEFs (Fig. 2A and B). These data suggest that Slc25a3 is vital for enhanced processing of Opa1 long isoforms in Drp1-KO MEFs. Supporting this notion, shRNA-mediated knockdown of Slc25a3 in Drp1-KO MEFs also increased the levels of L1 and L2 (Fig. S3). As a control, we knocked out the metalloprotease Oma1, which cleaves L1 and L2, in Drp1-KO MEFs. Results show that the loss of Oma1 also increases the levels of these long isoforms in Drp1Oma1-KO MEFs (Fig. S4). In addition, we found that the levels of two Drp1 receptors, Mff and Fis1, are not affected in the absence of Drp1 or Slc25a3 (Fig. 2A and B). Another Drp1 receptor, Mid49, exhibited increased levels in Drp1-KO MEFs, and these elevated levels were maintained in Drp1Slc25a3-KO MEFs (Fig. 1A and B).
Figure 2. OPA1 cleavage.

(A) Western blotting of WT, Drp1-KO, Slc25a3-KO, and Drp1Slc25a3-KO MEFs with the indicated antibodies. (B) Quantification of band intensity (average ± SD, n = 3). (C) Drp1-KO and Drp1Slc25a3-KO MEFs were transduced with lentiviruses carrying Opa1 L2- HA. The expression of L2-HA was induced for 4 h (0.1 μg/ml doxycycline). Western blotting was performed with the indicated antibodies. (D) Quantification of band intensity (average ± SD, n = 3). Significance was calculated using ANOVA with post-hoc Tukey in (B) and Student’s t-test in (D): *p<0.05, **p<0.01, ***p<0.001. See also Figures S3, S4, and S5.
If the cleavage of L1 and L2 by Oma1 is decreased, it is anticipated that levels of their cleaved products, S3 and S5, respectively, may be reduced. However, we did not observe such decreases in Drp1Slc25a3-KO MEFs (Fig. 2A and B) or Drp1-KO MEFs depleted for Slc25a3 (Fig. S3). It is possible that the loss or depletion of Slc25a3 may affect the stability of S3 and S5, making the data interpretation challenging. To unambiguously test the role of Slc25a3 in Opa1 processing, we expressed Opa1-L2-HA from an inducible Tet-On promoter in Drp1-KO and Drp1Slc25a3-KO MEFs, as we did previously 19. An induced expression of Opa1-L2-HA for a short time allowed us to directly assess the Oma1-mediated proteolytic cleavage of Opa1-L2 to Opa1-S5 19. We found that the loss of Slc25a3 significantly decreased the conversion of Opa1 L2-HA to S5-HA (Fig. 2C and D). These data suggest that Slc25a3 is crucial for the cleavage of Opa1 in Drp1-KO cells.
Furthermore, to ask if the transcription of Opa1-L1 and-L2 is altered in Drp1-KO and Drp1Slc25a3-KO MEFs, we examined their mRNA levels. In contrast to their decreased protein levels, the mRNA levels of both L1 and L2 were increased in Drp1-KO cells (Fig. S5). The increased mRNA levels might represent a compensatory response to their reduced protein abundance. Notably, in Drp1Slc25a3-KO cells, while L2 transcript levels remain elevated, L1 transcript levels have restored (Fig. S5).
The simultaneous loss of Slc25a3 and Drp1 produces megamitochondria and clusters mitochondrial DNA
Flickering-induced Opa1 processing safeguards mitochondria from excess fusion 19. To test whether decreased flickering inhibits this defense mechanism in Drp1Slc25a3-KO cells, we examined mitochondrial morphology using laser confocal immunofluorescence microscopy with antibodies to a matrix protein (pyruvate dehydrogenase, PDH) and an outer membrane protein (Tom20) 19. As reported 19,37, mitochondria became elongated and created spherical structures in Drp1-KO MEFs (Fig. 3A–C). In Slc25a3-KO MEFs, mitochondrial morphology appeared normal (Fig. 3A–C). Strikingly, in Drp1Slc25a3-KO MEFs, both the frequency and size of spherical mitochondria were dramatically increased compared to Drp1-KO MEFs (Fig. 3A–C). Although mitochondrial morphology seems unaffected in Slc25a3-KO MEFs when imaged by light microscopy, electron microscopy showed a remarkable change in the ultrastructure of mitochondria (Fig. 3D). Inner membrane cristae were greatly lost, and only short, infrequent cristae structures were found in Slc25a3-KO MEFs (Fig. 3D). In Drp1-KO MEFs, elongated mitochondria maintain normal cristae structures (Fig. 3D). These structural defects in the single KO MEFs were combined in the double-KO MEFs. We found elongated and enlarged mitochondria without the majority of cristae in Drp1Slc25a3-KO MEFs (Fig. 3D).
Figure 3. Mitochondrial morphology and cristae structure.

(A) WT, Drp1-KO, Slc25a3-KO, and Drp1Slc25a3-KO MEFs were subjected to laser confocal immunofluorescence microscopy with antibodies to PDH and Tom20. The boxed regions are magnified. Scale bar, 10 μm. (B) Quantification of mitochondrial morphology (average ± SD, n = 3 experiments). In each experiment, 10 cells were analyzed. (C) The size of spherical mitochondria (average ± SD, n = 7 for WT, 79 for Drp1-KO, 3 for Slc25a3-KO, and 211 for Drp1Slc25a3-KO). (D) Mitochondria in the same set of MEFs were analyzed by transmission electron microscopy. Scale bar, 1 μm. Significance was calculated using ANOVA with post-hoc Tukey in (C): ***p<0.001. See also Figure S6.
To determine how these defects in mitochondrial morphology affect the distribution of mitochondrial DNA (mtDNA) 38–40, we visualized mtDNA in WT, Drp1-KO, Slc25a3-KO, and Drp1Slc25a3-KO MEFs using a DNA-binding fluorescent dye (SYBR Green I) and MitoTracker 41. Live-cell imaging showed that mtDNA signals are present along mitochondrial tubules in WT and Slc25a3-KO MEFs (Fig. 4A and B). In Drp1-KO MEFs, mtDNA signals were found in spherical parts of mitochondria in addition to tubules. Strikingly, in Drp1Slc25a3-KO MEFs, mtDNA signals were almost completely lost in the tubular parts of mitochondria (Fig. 4A, Box 1), and the majority of mtDNA was accumulated in enlarged spherical parts of mitochondria (Fig. 4A, Box 2). In these spherical mitochondria, the area of mtDNA signal was increased (Fig. 4A, Box 2, and 4B). To further analyze mtDNA, we immunostained Drp1Slc25a3-KO MEFs using antibodies to DNA and Hsp60 (a matrix protein). Laser confocal microscopy showed that mtDNA molecules are clustered in spherical mitochondria (Fig. 4C). Therefore, these data suggest that Drp1 and Slc25a3 synergistically maintain mtDNA’s distribution. As a control, we knocked out the Opa1-cleaving protease Oma1 in Drp1-KO MEFs and analyzed mitochondrial morphology and mtDNA distribution. Similar to Drp1Slc25a3-KO MEFs, Drp1Oma1-KO MEFs showed enlarged mitochondria with clustered mDNA (Fig. S6).
Figure 4. mtDNA distribution.

(A) MEFs were stained with SYBR Green I and MitoTracker and viewed using live-cell imaging. Scale bar, 10 μm. (B) The size of mtDNA signal (average ± SD, n = 30). (C) Drp1Slc25a3-KO MEFs were analyzed by laser confocal immunofluorescence microscopy with antibodies to DNA and Hsp60. (D-J) WT, Drp1-KO, and Drp1Slc25a3-KO MEFs were transfected with a plasmid carrying mUNG (Y147A). (D) Seven days after transfection, cells were subjected to laser confocal immunofluorescence microscopy. Scale bar, 10 μm. (E, G, I) The percentage of cells with spherical mitochondria (average ± SD, n = 30). (F, H, J) The size of spherical mitochondria (average ± SD, n = ~20 [F], ~30 [H], and ~130 [J]). (K) Flickering frequency is quantified in the indicated cells (average ± SD, n = 3). Significance was calculated using ANOVA with post-hoc Tukey in (B and K) and Student’s t-test in (F, H, J): *p<0.05, ***p<0.001.
To ask if the accumulation of mtDNA is involved in forming enlarged mitochondrial spheres in Drp1Slc25a3-KO MEFs, we transfected WT, Drp1-KO, and Drp1Slc25a3-KO cells with a plasmid carrying a matrix-targeted uracil-N-glycosylase mutant, mUNG1(Y147A) (Fig. 4D) 42. WT uracil-N-glycosylase functions in DNA repair and removes misincorporated uracil from DNA. The mutation in the active site (Y147A) changes its specificity and enables this enzyme to remove normally incorporated thymine 42. The resulting single-strand gaps in mtDNA cause the depletion of mtDNA via degradation 42. Transfected cells with matrix-targeted mUNG1(Y147A) were cultured in the presence of uridine and pyruvate to support cell proliferation without mtDNA 42. Seven days after the transfection, cells were immunostained with antibodies to DNA and Hsp60. As expected, mUNG1(Y147A) greatly eliminated mtDNA in WT, Drp1-KO, and Drp1Slc25a3-KO MEFs (Fig. 4D). We found that the depletion of mtDNA significantly decreased the percentage of cells that contain spherical mitochondria and the size of spherical mitochondria in both Drp1-KO and Drp1Slc25a3-KO MEFs (Fig. 4E–J). In addition, we analyzed flickering in Drp1-KO and Drp1Slc25a3-KO cells carrying mUNG1 (Y147A) and found that flickering frequency was not affected by the removal of mtDNA (Fig. 4K). These data suggest that the clustering of mtDNA contributes to increases in mitochondrial size, but not flickering, in Drp1-KO and Drp1Slc25a3-KO cells, possibly by physically clogging and bulging mitochondrial tubules.
Impacts of the loss of Drp1 and Slc25a3 on oxidative phosphorylation and glycolysis
To test the effects of the loss of Slc25a3 on energy metabolism, we first measured mitochondrial oxygen consumption rates (OCRs) in WT, Drp1-KO, Slc25a3-KO, and Drp1Slc25a3-KO MEFs. We found that individual loss of Drp1 or Slc25a3 decreased OCRs, and their simultaneous loss further decreased OCRs at the basal and maximal levels (Fig. 5A and B). These additive effects suggest separate roles of Drp1 and Slc25a3 for mitochondrial respiration. Second, we analyzed glycolysis by measuring extracellular acidification rates (ECARs). In contrast to OCRs, basal ECARs were increased in Drp1-KO, Slc25a3-KO, and Drp1Slc25a3-KO MEFs (Fig. 5C and D). These changes may be a compensatory response to decreased mitochondrial respiration. Indeed, we found that glycolytic reserve was reduced in Drp1-KO, Slc25a3-KO, and Drp1Slc25a3-KO MEFs without gross changes in glycolytic capacity (Fig. 5D).
Figure 5. The role of copper transport by Slc25a3 for flickering and Opa1 processing.

(A) Mitochondrial respiration was analyzed by measuring OCRs in WT, Drp1-KO, Slc25a3-KO, and Drp1Slc25a3-KO MEFs. (B) The basal and maximal OCRs (average ± SD, n = 3). (C) Glycolysis was examined by measuring ECARs in the same set of MEFs. (D) The basal glycolysis, glycolytic capacity, and glycolytic reserve (average ± SD, n = 3). (E) Drp1-KO MEFs and Drp1Slc25a3-KO MEFs expressing WT and mutant forms of the mouse (m) and human (h) Slc25a3 were analyzed for flickering (average ± SD, n = 3 experiments). In each experiment, about 15 cells were analyzed. (F) Drp1Slc25a3-KO MEFs were incubated with 0.1 μM Cu-ATSM for 3 days and analyzed for flickering (average ± SD, n = 3 experiments). (G) Copper levels in mitochondria isolated from Drp1Slc25a3-KO MEFs treated with Cu-ATSM were measured using atomic absorption spectrometry. (average ± SD, n = 3 experiments). (H) Western blotting of Drp1-KO MEFs and Drp1Slc25a3-KO MEFs expressing the indicated WT and mutant mouse Slc25a3. (I) Quantification of band intensity (average ± SD, n = 3). Significance was calculated using ANOVA with post-hoc Tukey in (B, D, E, and I) and Student’s t-test in (F and G): *p<0.05, **p<0.01, ***p<0.001.
Slc25a3-mediated copper transport is critical for flickering, Opa1 processing, and mitochondrial morphology in Drp1-KO cells
Recent excellent studies have shown that Slc25a3 transports copper in addition to phosphate 32,33. To ask which ions are important for the role of Slc25a3 in flickering, we introduced mutations to Slc25a3, which block the transport of both copper and phosphate (H75A in mouse Slc25a3) or specifically phosphate (L175A) 33. We expressed WT Slc25a3 and the mutant versions in Drp1Slc25a3-KO MEFs and measured flickering. We found that Slc25a3 (H75A), which is defective in the transport of both copper and phosphate, failed to rescue decreased flickering in Drp1Slc25a3-KO MEFs (Fig. 5E). In contrast, Slc25a3 (L175A), which is defective in only phosphate transport, significantly increased flickering frequency in Drp1Slc25a3-KO MEFs, similar to WT Slc25a3 (Fig. 5E). To directly test the importance of copper transport for flickering, we treated Drp1Slc25a3-KO MEFs with a membrane-permeable copper-containing compound, diacetylbis(N(4)-methylthiosemicarbazonato)copper(II) (Cu-ATSM) 33. Noticeably, Cu-ATSM treatment effectively increased the flickering rate in Drp1Slc25a3-KO MEFs (Fig. 5F). We confirmed that Cu-ATSM increases the level of mitochondrial copper in Drp1Slc25a3-KO MEFs using atomic absorption spectrometry of isolated mitochondria (Fig. 5G). Therefore, a decrease in copper transport into mitochondria reduced the flickering rate in Drp1Slc25a3-KO MEFs. Consistent with the impact on flickering, WT and Slc25a3 (L175A), but not Slc25a3 (H75A), rescued Opa1 processing and decreased levels of L1 and L2 in Drp1Slc25a3-KO MEFs (Fig. 5H and I). Unlike Opa1, the decreased levels of Mfn1 and Mfn2 in Drp1Slc25a3-KO MEFs were not restored by either WT Slc25a3 or its mutants (Fig. 5H and I). We confirmed the effects of the Slc25a3 mutants on Opa1 processing by utilizing the Opa1-L2-HA construct (Fig. S7A and B). These data suggest that copper transport by Slc25a3 plays a crucial role in flickering and Opa1 processing in Drp1-KO cells.
To determine whether the copper transport activity of Slc25a3 is required for mitochondrial morphology in the absence of Drp1, we performed laser confocal immunofluorescence microscopy with antibodies to PDH and Tom20 in Drp1Slc25a3-KO MEFs expressing WT or the mutants of Slc25a3. Consistent with the effects on flickering and Opa1 processing, the expression of WT Slc25a3 and Slc25a3 (L175A) reduced the percentage of cells that contain giant spherical mitochondria in Drp1Slc25a3-KO MEFs (Fig. 6A and B) and the size of individual spherical mitochondria (Fig. 6C). Conversely, Slc25a3 (H75A) barely rescued mitochondrial morphology in Drp1Slc25a3-KO MEFs (Fig. 6A–C). These data suggest that the copper transport by Slc25a3 is important for mitochondrial morphology in the absence of Drp1. In addition, we found that the L175A mutant rescues the cristae structure in Slc25a3-KO cells while the H75A mutant does not (Fig. 6D). Furthermore, the L175A mutant fully rescued both basal and maximal OCRs in Slc25a3-KO cells, but the H75A mutant did not rescue basal OCRs and only partially increased maximal OCRs (Fig. 6E–G). These data suggest that copper transport activity by Slc25a3 is crucial for maintaining cristae structure and mitochondrial respiration.
Figure 6. The role of copper transport by Slc25a3 for mitochondrial morphology, cristae structure, and respiration.

(A) Drp1-KO MEFs and Drp1Slc25a3-KO MEFs expressing the mouse (m) and human (h) Slc25a3 were subjected to laser confocal immunofluorescence microscopy with antibodies to PDH and Tom20. The boxed regions are magnified. Scale bar, 10 μm. (B) Quantification of cells with spherical mitochondria (average ± SD, n = 3 experiments). In each experiment, 10 cells were analyzed. (C) The size of spherical mitochondria (average ± SD, n >70). (D) Mitochondria in the indicated MEFs were analyzed by transmission electron microscopy. Scale bar, 500 nm. (E) Mitochondrial respiration was analyzed by measuring OCRs in Slc25a3-KO MEFs expressing the indicated constructs. (F and G) The basal OCRs (F) and maximal OCRs (G) (average ± SD, n = 3). Significance was calculated using ANOVA with post-hoc Tukey in (C, F, and G): *p<0.05, ***p<0.001.
Copper-binding enzymes, superoxide dismutase 1 and cytochrome c oxidase, control flickering in opposite ways
Mitochondria contain two copper-containing enzymes, superoxide dismutase 1 (SOD1) and cytochrome c oxidase 43. To test whether these enzymes play a role in flickering, we first treated Drp1-KO MEFs with a SOD1 inhibitor (LCS-1) and examined flickering. We found that LCS-1 elevated the flickering frequency in Drp1-KO MEFs (Fig. 7A) and, notably, in WT MEFs as well (Fig. 7B). In line with these observations, LCS-1 also decreased the levels of Opa1-L2-HA in Drp1-KO MEFs (Fig. 7C and D) and reduced the formation of enlarged mitochondria (Fig. 7E–G). Conversely, the overexpression of SOD1 resulted in a decrease in flickering frequency (Fig. 7H). These findings indicate that SOD1 acts as a negative regulator of mitochondrial flickering, Opa1 processing, and the formation of spherical mitochondria. Second, we tested the role of cytochrome c oxidase using its inhibitor NaN3. In contrast to the SOD1 inhibitor LCS-1, NaN3 treatment reduced flickering frequency in Drp1-KO MEFs (Fig. 7I). Therefore, it appears that copper affects flickering via two enzymes in an antagonistic way. SOD1 suppresses flickering while cytochrome c oxidase promotes it.
Figure 7. The effects of SOD1 and cytochrome c oxidase on flickering, and the impact of disease-associated Slc25a3 mutants on Opa1 processing and copper transport.

(A and B) Drp1-KO MEFs (A) and WT MEFs (B) were treated with 0.1 mM LCS-1 for 1 h and analyzed for flickering (average ± SD, n = 3 experiments). (C) Drp1-KO MEFs carrying Opa1 L2-HA were treated with LCS-1. The expression of L2-HA was induced for 4 h (0.1 μg/ml doxycycline). Western blotting was performed with the indicated antibodies. (D) Quantification of band intensity (average ± SD, n = 3). (E) Drp1-KO MEFs were treated with LCS-1 and then analyzed by laser confocal immunofluorescence microscopy with antibodies to DNA and Hsp60. Scale bar, 10 μm. (F) The percentage of cells with spherical mitochondria (average ± SD, n = 30). (G) The size of spherical mitochondria (average ± SD, n = 66 for DMSO and 98 for LCS-1). (H) Drp1-KO MEFs were transduced with lentiviruses carrying SOD1-FLAG and analyzed for flickering (average ± SD, n = 3 experiments). (I) Drp1-KO MEFs were treated with 1 mM NaN3 for 1 h and analyzed for flickering (average ± SD, n = 3 experiments). (J) Western blotting of Drp1-KO MEFs and Drp1Slc25a3-KO MEFs expressing the indicated human Slc25a3 constructs. (K) Quantification of band intensity (average ± SD, n = 3). (L) Copper levels in mitochondria isolated from WT MEFs and Drp1Slc25a3-KO MEFs expressing either WT Slc25a3 or the G72E mutant were measured using atomic absorption spectrometry (average ± SD, n = 3 experiments). Significance was calculated using Student’s t-test in (A, B, D, G, H, and I) and ANOVA with post-hoc Tukey in (K and L): *p<0.05, **p<0.01, ***p<0.001. See also Figure S7.
The pathogenic mutation G72E inhibits the copper transport activity of human Slc25a3
Defects in human Slc25a3 lead to fatal mitochondrial defects with symptoms of lactic acidosis, hypertrophic cardiomyopathy, and neonatal hypotonia 30,31,44–46. To test the impact of disease-associated mutations on flickering, we cloned human Slc25a3 and introduced the disease-associated G72E mutation in the first transmembrane α-helix 31. WT human Slc25a3 rescued Opa1 cleavage in Drp1Slc25a3-KO MEFs, while Slc25a3 (G72E) failed to do so, although their expression levels were similar (Fig. 7J and K). This result concerning Opa1 processing was confirmed by employing Opa1-L2-HA (Fig. S7C and D). Similarly, the G72E mutation blocked the ability of Slc25a3 to rescue flickering frequency (Fig. 5E) and mitochondrial morphology (Fig. 6A–C) in Drp1Slc25a3-KO MEFs. In addition, we found that the G72E mutant failed to restore cristae morphology and mitochondrial respiration in Slc25a3-KO MEFs (Fig. 6D–G). Finally, atomic absorption spectrometry of isolated mitochondria showed that Slc25a3 (G72E) could not replenish mitochondrial copper levels in Drp1Slc25a3-KO MEFs (Fig. 7L). These findings suggest that the pathogenic G72E mutation impairs the copper transport activity of human Slc25a3, thereby affecting its critical role in maintaining mitochondrial structure and function.
Discussion
The mitochondrial defense mechanism, MitoSafe, regulates a balance between mitochondrial fusion and division when mitochondria are hyper-fused. MitoSafe decreases levels of outer membrane fusion proteins, Mfn1 and Mfn2, by the ubiquitin E3 ligase Parkin and the protein kinase PINK1, and proteolytically inactivates Opa1 by the metalloprotease Oma1, which is activated by the flickering of the mitochondrial membrane potential 19–21. In the current study, using a live-cell imaging-based gene knockdown screen, we report that the carrier protein Slc25a3 in the mitochondrial inner membrane plays a critical role in flickering in hyper-fused mitochondria. Slc25a3 transports phosphate and copper ions across the inner membrane using the proton gradient 32,33. We found that the transport of copper, rather than phosphate, by Slc25a3 is critical for flickering in MitoSafe.
Copper is an essential heavy metal but highly toxic in cells 43. The concentration of free copper in cells is extremely low (estimated < 1 attomolar) and tightly regulated 47. Excess or reduced levels of copper cause human diseases such as Wilson’s disease (due to copper overload) and Menkes disease (due to copper deficiency), both of which are associated with impaired mitochondrial function 48. In mitochondria, copper is required for complex IV (cytochrome c oxidase) and an antioxidant enzyme, superoxide dismutase 1 (SOD1). Cytochrome c oxidase is a multisubunit enzyme that contains copper-binding subunits such as Cox1 and Cox2 for the catalytic activity. SOD1 converts superoxide to hydrogen peroxide in the cytosol and mitochondria, binds to copper, and uses it as a cofactor along with zinc. It has been shown that the loss of Slc25a3 alters levels of copper in the matrix, cytochrome c oxidase function, and SOD1 activities in cells 32. As downstream effectors of copper, we found that cytochrome c oxidase and SOD1 function in flickering. Interestingly, they antagonistically control flickering. While cytochrome c oxidase stimulates flickering, SOD1 suppresses it.
How do cytochrome c oxidase and SOD1 regulate flickering? Cytochrome c oxidase facilitates the generation of membrane potential across the inner membrane as part of the electron transport chain; thus, inhibition of cytochrome c oxidase may decrease the membrane potential, which could influence flickering in Drp1-KO cells. This model aligns with the observation that the complex V inhibitor oligomycin, which induces the hyperpolarization of the inner membrane, stimulates flickering 19. However, since we observed normal levels of membrane potential in Drp1Slc25a3-KO cells, this possibility is unlikely. Alternatively, since SOD1 counteracts mitochondrial ROS, cytochrome c oxidase and SOD1 might affect ROS levels in Drp1-KO cells in opposite ways. The role of cytochrome c oxidase in ROS production is highly context-dependent, and its inhibition can lead to decreased ROS production 49–51. In scenarios where the electron transport chain is dysfunctional and hyperactive, an increased flow of electrons can enhance the probability of electron leakage and subsequent ROS formation, particularly at complexes I and III, known sites of superoxide generation 49–51. Inhibiting cytochrome c oxidase could slow the electron flow, potentially reducing the chance of electron leakage and the formation of ROS. This model is consistent with our observations that Drp1-KO MEFs maintain relatively normal levels of basal OCRs but exhibit decreased respiratory capacity, suggesting that the mitochondria are dysfunctional and excessively utilize spare respiratory capacity. Moreover, Drp1-KO cells display elevated levels of oxidative damage 52. Therefore, it is conceivable that SOD1 and cytochrome c oxidase impact ROS in this specific mitochondrial condition in Drp1-KO cells. This model predicts that a yet-to-be-identified mitochondrial flickering pore is likely regulated by ROS. An important future direction is to elucidate its molecular identity.
Deficiency in human Slc25a3 leads to mitochondrial disorders, manifested by cardiomyopathy, hypotonia, and lactic acidosis 30,31,44,45. Our findings indicate that the pathogenic mutation G72E impairs Slc25a3’s role in mitochondrial copper transport, cristae structure, and mitochondrial respiration. Strikingly, our analysis of the mouse H75A and L175A mutants demonstrates that defects are more profoundly affected by the absence of copper transport than by phosphate transport. These data suggest that diseases associated with Slc25a3 deficiencies may primarily arise from disruptions in copper transport rather than phosphate transport. We propose that targeting mitochondrial copper bioavailability with copper-containing compounds, such as Cu-ATSM, could offer an effective treatment for Slc25a3-related diseases.
Limitation of the study
There are several key questions that remain to be addressed. For example, identifying a specific type of ROS that controls flickering would be important. We are also interested in deciphering the dynamics of copper during mitochondrial fusion, division, and flickering. Furthermore, investigating the role of copper-dependent mitochondrial defense mechanisms in vivo would provide further insight into the physiological regulation of mitochondrial dynamics.
STAR Methods
Resource Availability
Lead Contact
Further information and requests for resources and reagents should be directed to and will be fulfilled by the Lead Contact, Hiromi Sesaki (hsesaki@jhmi.edu).
Materials Availability
All unique reagents generated in this study are available from the lead contact with a completed materials transfer agreement.
Data and Code Availability
All data reported in this paper will be shared by the lead contact upon request.
This paper does not report original code.
Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request.
Experimental Model and Study Participant Details
Cells
MEFs were cultured in Iscove’s modified Dulbecco’s medium containing 10% fetal bovine serum 54. HEK293T cells were cultured in Dulbecco’s Modified Eagle Medium containing 10% fetal bovine serum. Slc25a3-KO, Drp1Slc25a3-KO, Oma1-KO, and Drp1Oma1-KO MEFs and Drp1-KO HEK293T cells were generated using a GeneArt CRISPR Nuclease (OFP Reporter) Vector Kit (A21174; Thermo Fisher Scientific) following the manufacturer’s instructions. The target gRNA sequence was 5′-TTGGTGGGGTCTTAAGTTGT-3′ for Slc25a3-KO MEFs, 5′-GAGTGAATAACCTGGCCAAC-3′ for Oma1-KO MEFs, and 5′-GCCTGTAGGTGATCAACCTA-3′ for Drp1-KO HEK293T cells. Drp1Slc25a3-KO and Drp1Oma1-KO MEFs were generated by introducing the gRNA into Drp1-KO MEFs 19. Based on OFP fluorescent signal, transfected cells were sorted in 96-well plates as single cells at the Johns Hopkins School of Medicine Flow Cytometry Core 56. The loss of Slc25a3, Oma1, and Drp1 was confirmed by Western blotting.
Method Details
Flickering assay
To observe flickering in MEFs, cells were plated at a density of 20,000 or 40,000 cells/well in 8-well chambered coverglasses and cultured for 24 h 19. Cells were then stained with 5 nM TMRE in Iscove’s modified Dulbecco’s medium containing 10% FBS at 37°C with 5% CO2 for 15 min and examined using a Zeiss LSM800 GaAsP confocal microscope with a 40× objective lens at 37°C with 5% CO2. To minimize potential phototoxicity, images were obtained at 10-s intervals for 30 min at a single focal plane 19. In the esiRNA-based screen, cells were plated at a density of 15,000 cells/well for Drp1-KO MEFs and 2000 cells/well for Drp1-KO HEK293T cells in 8-well chambered coverglasses and cultured for 24 h. Cells were transfected with esiRNAs (Eupheria Biotech) at 100 ng/well using Lipofectamine RNAiMAX (Invitrogen). After 3 days, the culture medium was replaced with fresh medium and cells were transfected with esiRNAs again. Cells were cultured for an additional 2 days. Cells were stained with 5 nM TMRE and were viewed as described above. To quantify flickering, we counted cells that showed flickering at least once during the observation period of 30 min. We repeated the experiments three times, except for the esiRNA screen, in which we repeated the experiment twice. In each experiment, we analyzed more than 30 cells.
Plasmids
To generate shRNA plasmids, the following target sequences were cloned into pLKO.1. Scramble: CCTAAGGTTAAGTCGCCCTCGctcgagCGAGGGCGACTTAACCTTAGG, Ccdc51: GAAGAGAAGAGGCTCCGAATActcgagTATTCGGAGCCTCTTCTCTTC and GTGAACAGGCTTCTAGCTATTctcgagAATAGCTAGAAGCCTGTTCAC, Mpv17: CCCACGAATAGACACGCATTTctcgagAAATGCGTGTCTATTCGTGGG and GCTGGATCACTGATGGGCGTActcgagTACGCCCATCAGTGATCCAGC, Slc25a3: GCAACATACTTGGTGAGGAAActcgagTTTCCTCACCAAGTATGTTGC and CGACTCTGTGAAGGTCTACTTctcgagAAGTAGACCTTCACAGAGTCG, Slc25a40: CATCCACCTCTAGATCATAATctcgagATTATGATCTAGAGGTGGATG.
The tetracycline-inducible lentiviral vector pInducer20 carrying HA-tagged Opa1 (L2-HA) was previously generated 19. To induce the expression of L2-HA, MEFs were incubated with 0.1 μg/ml doxycycline for 4 h 19. Mouse Slc25a3 was PCR-amplified from total cDNAs of WT MEFs using the following primers, 5’-TAGGGATCCGCCACCATGTTCTCGTCCGTAGCGCACC-3’ and 5’-TAGGCGGCCGCCTACTCAGTTAACCCAAGCTTCTTCTTCAGAG-3’, and cloned into the BamHI and Not I sites of the pHR-Sin plasmid. Codon-optimized human Slc25a3 was PCR-amplified from the pDONR221_SLC25A3 plasmid (132040, Addgene) using primers 5’-AGACTGAGTCGCCCGGGGGGGATCCGCCACCATGTTCAGCTCCGTGGCACAC-3’ and 5’-CAGGTCGACTCTAGAGTCGCGGCCGCCTACTGTGTCAGGCCCAGCTTC-3’, and cloned into the pHR-Sin plasmid digested with BamHI and Not I using Gibson Assembly (E2611, New England Biolabs). The mutant Slc25a3 plasmids were generated by replacing the nucleotides (223C>G and 224A>C for mouse H75A, 523C>G and 524T>C for mouse L175A, and 215G>A and 216C>G for human G72E) in the wildtype Slc25a3 plasmids. Mutations are underlined in the sequences below. First, two partial fragments of Slc25a3 were PCR-amplified from the wildtype Slc25a3 plasmids using the following primers. H75A: 5’-TAGGGATCCGCCACCATGTTCTCGTCCGTAGCGCACC-3’ and 5’-CAACAGCAGTGGCTGTCAGCCCAC-3’. and 5’-GTGGGCTGACAGCCACTGCTGTTG-3’ and 5’-TAGGCGGCCGCCTACTCAGTTAACCCAAGCTTCTTCTTCAGAG-3’. L175A: 5’-TAGGGATCCGCCACCATGTTCTCGTCCGTAGCGCACC-3’ and 5’-CATAGGAGCCGCGGCAATGTCAG-3’, and 5’-CTGACATTGCCGCGGCTCCTATG-3’ and 5’-TAGGCGGCCGCCTACTCAGTTAACCCAAGCTTCTTCTTCAGAG-3’. G72E: 5’-AGACTGAGTCGCCCGGGGGGGATCCGCCACCATGTTCAGCTCCGTGGCACAC-3’ and 5’-TGATCTCGCCCAGGCCGCACAG-3’, and 5’-GGCCTGGGCGAGATCATCTCCTG-3’ and 5’-CAGGTCGACTCTAGAGTCGCGGCCGCCTACTGTGTCAGGCCCAGCTTC-3’. Second, full-length Slc25a3 mutants were PCR-amplified from the two products of the first PCR using the following primers. H75A and L175A: 5’-TAGGGATCCGCCACCATGTTCTCGTCCGTAGCGCACC-3’ and 5’-TAGGCGGCCGCCTACTCAGTTAACCCAAGCTTCTTCTTCAGAG-3’, and G72E: 5’-AGACTGAGTCGCCCGGGGGGGATCCGCCACCATGTTCAGCTCCGTGGCACAC-3’ and 5’-CAGGTCGACTCTAGAGTCGCGGCCGCCTACTGTGTCAGGCCCAGCTTC-3’. The PCR products were cloned into the BamHI and Not I sites of the pHR-Sin plasmid.
Lentivirus
HEK293T cells were seeded at 1.5 × 106 cells in a 10-cm dish and cultured for 24 h. To produce lentiviruses, 3 μg of the pHR-Sin plasmid carrying Slc25a3, the pInducer20 plasmid carrying L2-HA, or pLKO. 1 carrying shRNAs was co-transfected into HEK293T cells along with 3 μg of pHR-CMV8.2ΔR (for pHR-Sin and pLKO.1) or pHR-CMV8.9ΔR (for pInducer20), and 0.3 μg of pCMV-VSVG using Lipofectamine 2000 (Invitrogen) 19,41. After 20–22 h, the culture medium was replaced with fresh medium. After an additional 24 h, the culture medium containing the released viruses was collected. Lentiviruses carrying L2-HA were concentrated using Lenti-X Concentrator (Clontech). For lentiviral transduction, MEFs were seeded at 8 × 104 cells/well in a 6-well plate and cultured for 24 h. Cells were then incubated with lentivirus in the cell culture medium containing 10% FBS and 8 μg/ml polybrene for 24 h.
Western blotting
Cells were harvested and lysed in RIPA buffer (9806S, Cell Signaling Technology) supplemented with cOmplete, Mini, EDTA-free Protease Inhibitor Cocktail (11836170001, Roche) on ice 19. The lysates were centrifuged at 16,000 g for 10 min at 4°C, and the supernatants were collected. Proteins were separated by SDS-PAGE and transferred onto Immobilon-FL Transfer Membrane (Millipore). The membranes were blocked in PBS-T (PBS containing 0.05% Tween 20) containing 3% BSA at room temperature for 1 h and then incubated with primary antibodies in PBS-T containing 3% BSA at 4°C overnight. The antibodies used were Opa1 (1:1,000 dilution, 612607; BD Biosciences), Drp1 (1:2,000 dilution, 611113; BD Biosciences), Slc25a3 (1:1,000 dilution, H00005250-B02P; Novus Biologicals) (Fig. 2A, 2C, 5H, S3A, and S7A), Slc25a3 (1:1,000 dilution, 10420-1-AP; Proteintech) (Fig. 7K and S7C), mitofusin 1 (1:1,000 dilution, ab126575; Abcam), mitofusin 2 (1:1,000 dilution, ab57602; Abcam), α-tubulin (1:1,000 dilution, 2125; Cell Signaling Technology), HA (1:2,000 dilution, NB600-362; Novus Biologicals), Mff (a kind gift from Dr. Alexander M. van der Bliek, UCLA)53, Fis1 (1:1,000, 10956-1-AP, Proteintech), MiD49 (1:1,000, 16413-1-AP, Proteintech), and Oma1 (1:1,000, sc-515788, Santa Cruz Biotechnology). The membranes were washed three times in PBS-T, followed by incubation with appropriate secondary antibodies at room temperature for 1 h. After washing the membranes three times in PBS-T, fluorescence signals were detected using a Typhoon laser-scanner platform (Amersham). To quantify band intensity, we magnified Western blot images using the NIH FIJI program and selected individual bands using a box tool. We also selected a background area in the vicinity of the bands using a box tool of the same dimensions. Band intensity was determined after subtracting the background intensity. We also developed a method to analyze individual isoforms using a line scan tool in FIJI when they are closely located 57.
Measurements of the mitochondrial membrane potential
The mitochondrial membrane potential was measured using Cell Meter NIR Mitochondrial Membrane Potential Assay Kit (AAT Bioquest) following the manufacturer’s instructions 19. MEFs were first incubated with MitoLite NIR at 37°C and 5% CO2 for 30 min. MEFs were resuspended in 1 ml of the supplied assay buffer and filtered through a 70 μm cell strainer (22363548, Fisher Scientific). The fluorescence intensity was measured using a FACSCalibur (BD Biosciences).
qPCR
Cells were plated on 12-well plates. Total RNA was purified from the cells using the RNeasy Mini Kit (74106, Qiagen) and reverse-transcribed using the ReadyScript cDNA Synthesis Mix (RDRT, Sigma-Aldrich). PCR was performed using the QuantStudio 3 Real-Time PCR System (Applied Biosystems) with the PowerUp SYBR Green Master Mix (A25741, Thermo Fisher Scientific). GAPDH was used as the reference gene. The following primers were used: 5’-TTCTTCACTGCAGGTTCACCTG-3’ and 5’-AGAGAATGAGCTCACCAAGCAG-3’ for Opa1-L1, 5’-TTCTTCACTGCAGGTTCACCTG-3’ and 5’-TTCTTTGTCTGACACCTTCCTGTAA-3’ for Opa1-L2, and 5’-AAGGTCATCCCAGAGCTGAA-3’ and 5’-CTGCTTCACCACCTTCTTGA-3’ for GAPDH.
Immunofluorescence microscopy
MEFs were fixed in pre-warmed (37°C) PBS containing 4% paraformaldehyde for 20 min, washed three times in PBS, permeabilized with PBS containing 0.1% Triton X-100 for 8 min, washed again three times in PBS, and blocked in PBS containing 0.5% BSA at room temperature for 1 h 19,58. Cells were then incubated with anti-PDH antibody (1:300 in PBS containing 0.5% BSA, ab110333; Abcam), anti-Tom20 antibody (1:300, sc-11415; Santa Cruz Biotechnology), anti-DNA antibody (1:200, 690014S; PROGEN) and anti-Hsp60 antibody (1:400, 12165; Cell Signaling Technology) at 4°C overnight. Cells were washed three times in PBS and incubated with appropriate secondary antibodies at room temperature for 1 h, including Alexa 488-conjugated anti-mouse IgG (1:400 in PBS, A21202, Thermo Fisher Scientific), Alexa 568-conjugated anti-rabbit IgG (1:400, A10042, Thermo Fisher Scientific), Alexa 568-conjugated anti-mouse IgM (1:400, A21043, Thermo Fisher Scientific) and Alexa 647-conjugated anti-rabbit IgG (1:400, A31573, Thermo Fisher Scientific). Cells were washed three times in PBS. The samples were observed using an LSM800 GaAsP laser scanning confocal microscope 41,59. The size of enlarged spherical mitochondria was quantified using the NIH FIJI program (Fig. 3C, 4F, 4H, 4J, 6C, and S6C). To select such mitochondria, images displaying PDH or Hsp60 were first binarized. Then, the circularity parameter was set between 0.5 and 1.0, and a size cutoff threshold of greater than 2 μm2 was applied to ensure the selection of appropriately shaped mitochondria. The area of the selected mitochondria was then determined.
Electron microscopy
MEFs were fixed by 2% glutaraldehyde, 3 mM CaCl2 and 0.1 M cacodylate buffer, pH 7.4, for 1 h. After washes, samples were post-fixed in 1% OsO4, 1%. potassium hexacyanoferrate, and 0.1 M cacodylate, pH 7.4, for 1 h on ice. After washes in water, samples were incubated in 2% uranyl acetate for 30 min on ice. After dehydration using 50, 70, 90, and 100% ethanol, samples were embedded in EPON resin. Ultrathin sections were obtained using a Reichert-Jung ultracut E, stained with 2% uranyl acetate and lead citrate, and viewed using a transmission electron microscope (H-7600; Hitachi) equipped with a dual CCD camera (Advanced Microscopy Techniques).
Live-cell imaging for mitochondrial DNA
MEFs were plated at 10,000 cells/well in 8-well chambered coverglasses and cultured for 24 h. Cells were incubated with 200 nM MitoTracker Red CMXRos (M7512, Thermo Fisher Scientific) and SYBR Green I Nucleic Acid Gel Stain (1:100,000 dilution, S7563, Thermo Fisher Scientific) in Iscove’s modified Dulbecco’s medium containing 10% FBS at 37°C with 5% CO2 for 15 min. The samples were observed using an LSM800 GaAsP laser scanning confocal microscope at 37°C with 5% CO2.
Mitochondrial respiration
Mitochondrial OCRs were measured using an XF96 Extracellular Flux Analyzer (Seahorse Bioscience) 19,41. MEFs were seeded at 5,000 cells/well in an XF 96-well culture microplate and cultured for 24 h. MEFs were washed twice in XF base medium supplemented with 25 mM glucose, 4 mM L-glutamine, and 1 mM sodium pyruvate, and then cells were incubated at 37°C in a CO2-free incubator for 1 h. OCR measurement was performed according to the manufacturer’s instructions. Baseline OCR was recorded three times, and then, 1.6 μg/ml oligomycin, 1 μM FCCP, and 0.5 μM rotenone/antimycin A were sequentially injected into each well.
Glycolysis
Glycolysis was measured by analyzing the extracellular acidification rate (ECAR) using an XF96 Extracellular Flux Analyzer (Seahorse Bioscience). Cells were seeded at 5000 cells/well in an XF 96-well culture microplate and cultured for 24 h. Cells were washed twice in XF base medium supplemented with 4 mM L-glutamine, then incubated at 37°C in a CO2-free incubator for 1 h. ECAR measurement was performed according to the manufacturer’s instructions. Non-glycolytic acidification was recorded three times, and then 10 mM glucose, 1.6 μg/ml oligomycin, and 50 mM 2-deoxyglucose were sequentially injected into each well.
Copper measurements
MEFs were cultured in ten 10-cm dishes. The plates were washed with ice-cold PBS and then placed on ice. The cells were suspended in 1 ml/dish of homogenization buffer [10 mM HEPES-KOH buffer (pH 7.4) containing 0.22 M mannitol, 0.07 M sucrose, and protease inhibitors (11836170001, Roche)] and collected using a cell scraper. The cells were then washed twice for 5 minutes each at 4°C. To isolate the mitochondria, cells were homogenized with 10 strokes using a syringe fitted with a 27-gauge needle in the same buffer. The postnuclear supernatant was obtained by centrifugation at 800 g for 10 minutes at 4 °C, followed by an additional 5 minutes. The supernatant was further separated into a mitochondria-enriched pellet by centrifugation at 8,000 g for 10 minutes at 4°C. The pellet was washed in the homogenization buffer and centrifuged again at 8,000 g for 10 minutes at 4°C. The mitochondria-enriched pellet was resuspended in 50 μl of 20 mM HEPES buffer, pH 7.4, containing 1 mM EDTA, 1 mM EGTA, 8.5% sucrose, 0.1% NP-40, 0.2% Triton X-100, 1 mM 4-(2-aminoethyl)benzenesulfonyl fluoride hydrochloride (AEBSF), and protease inhibitors (11836170001, Roche) and stored at −80°C. 30 μl of the mitochondria lysates were mixed with 55 μl of HNO3 (Trace metal free, A509P212, Fisher Chemical) in metal-free 1.5 ml tubes (616201, Greiner Bio-One) and incubated at 70°C for 4 hours. Samples were mixed with 115 μl of UltraPure Distilled water (10977, Invitrogen) and stored at 4°C. Copper amounts in 20 μl of samples were determined using atomic absorption spectrometry (PinAAcle 900T, Perkin Elmer) with copper standards (5 ppb – 50 ppb, SC194-500, Fisher Scientific) and normalized relative to protein concentration using the Syngistix AA software version 3.1 for measurement and analysis.
Quantification and statistical analysis
All statistical analyses were performed using Prism (GraphPad). All of the statistical details are described in each figure legend.
Supplementary Material
Key resources table
| REAGENT or RESOURCE | SOURCE | IDENTIFIER |
|---|---|---|
| Antibodies | ||
| Mouse monoclonal anti-Opa1 | BD Biosciences | Cat# 612607; RRID: AB_399889 |
| Mouse monoclonal anti-Drp1 | BD Biosciences | Cat# 611113; RRID: AB_398424 |
| Mouse polyclonal anti-Slc25a3 | Novus Biologicals | Cat# H00005250-B02P; RRID: AB_2270489 |
| Rabbit polyclonal anti-Slc25a3 | Proteintech | Cat# 10420-1-AP; RRID: AB_2877735 |
| Mouse monoclonal anti-mitofusin 1 | Abcam | Cat# ab126575; RRID: AB_11141234 |
| Mouse monoclonal anti-mitofusin 2+ mitofusin 1 | Abcam | Cat# ab57602; RRID: AB_2142624 |
| Rabbit monoclonal anti-α-Tubulin | Cell Signaling Technology | Cat# 2125; RRID: AB_2619646 |
| Goat polyclonal anti-HA | Novus Biologicals | Cat# NB600-362; RRID: AB_10124937 |
| Rabbit polyclonal anti-Mff | Gandre-Babbe and van der Bliek53 | N/A |
| Rabbit polyclonal anti-Fis1 | Proteintech | Cat# 10956-1-AP; RRID: AB_2102532 |
| Rabbit polyclonal anti-Mid49 | Proteintech | Cat# 16413-1-AP; RRID: AB_2714217 |
| Mouse monoclonal anti-Oma1 | Santa Cruz Biotechnology | Cat# sc-515788; RRID: AB_2905488 |
| Mouse monoclonal anti-PDH | Abcam | Cat# ab110333; RRID: AB_10862029 |
| Rabbit polyclonal anti-Tom20 | Santa Cruz Biotechnology | Cat# sc-11415; RRID: AB_2207533 |
| Mouse monoclonal anti-DNA | PROGEN | Cat# 690014S; RRID: AB_2750935 |
| Rabbit monoclonal anti-Hsp60 | Cell Signaling Technology | Cat# 12165; RRID: AB_2636980 |
| Alexa 488-conjugated anti-mouse IgG | Thermo Fisher Scientific | Cat# A21202; RRID: AB_141607 |
| Alexa 488-conjugated anti-rabbit IgG | Thermo Fisher Scientific | Cat# A21206; RRID: AB_2535792 |
| Alexa 488-conjugated anti-goat IgG | Thermo Fisher Scientific | Cat# A21467; RRID: AB_2535870 |
| Alexa 568-conjugated anti-rabbit IgG | Thermo Fisher Scientific | Cat# A10042; RRID: AB_2534017 |
| Alexa 568-conjugated anti-mouse IgM | Thermo Fisher Scientific | Cat# A21043; RRID: AB_2535712 |
| Alexa 647-conjugated anti-rabbit IgG | Thermo Fisher Scientific | Cat# A31573; RRID: AB_2536183 |
| Chemicals, peptides, and recombinant proteins | ||
| Iscove’s modified Dulbecco’s medium (IMDM) | Gibco | Cat# 12440-053 |
| Dulbecco’s Modified Eagle’s Medium (DMEM) – high glucose | Sigma-Aldrich | Cat# D5796 |
| Fetal Bovine Serum (FBS) | Corning | Cat# 35-010-CV |
| Tetramethylrhodamine, ethyl ester (TMRE) | Thermo Fisher Scientific | Cat# T669 |
| Lipofectamine 2000 Transfection Reagent | Thermo Fisher Scientific | Cat# 11668019 |
| Lipofectamine RNAiMAX Transfection Reagent | Thermo Fisher Scientific | Cat# 13778150 |
| Phusion High-Fidelity DNA Polymerase | NEB | Cat# M0530 |
| Quick Ligation Kit | NEB | Cat# M2200 |
| Gibson Assembly Master Mix | NEB | Cat# E2611 |
| Lenti-X Concentrator | Takara Bio | Cat# 631231 |
| Polybrene Infection/Transfection Reagent | MilliporeSigma | Cat# TR-1003-G |
| RIPA Buffer (10X) | Cell Signaling Technology | Cat# 9806S |
| cOmplete, Mini, EDTA-free Protease Inhibitor Cocktail | Roche | Cat# 11836170001 |
| 4–20% Criterion TGX Precast Gels | Bio-Rad Laboratories | Cat# 5671095 |
| 7.5% Criterion TGX Precast Gels | Bio-Rad Laboratories | Cat# 5671025 |
| Sodium cacodylate, trihydrate | Electron Microscopy Sciences | Cat# 12300 |
| 25% glutaraldehyde | Electron Microscopy Sciences | Cat# 16220 |
| 4% Osmium Tetroxide (OsO4) | Electron Microscopy Sciences | Cat# 19190 |
| Potassium hexacyanoferrate(II), trihydrate | Sigma-Aldrich | Cat# P3289 |
| Uranyl acetate | Ted Pella, Inc. | Cat# 19481 |
| EMBED 812 RESIN | Electron Microscopy Sciences | Cat# 14900 |
| DDSA | Electron Microscopy Sciences | Cat# 13710 |
| NMA | Electron Microscopy Sciences | Cat# 19000 |
| DMP-30 | Electron Microscopy Sciences | Cat# 13600 |
| MitoTracker Red CMXRos | Thermo Fisher Scientific | Cat# M7512 |
| SYBR Green I Nucleic Acid Gel Stain | Thermo Fisher Scientific | Cat# S7563 |
| Seahorse XF Base Medium Minimal DMEM | Agilent | Cat# 102353-100 |
| L-Glutamine (200 mM) | Gibco | Cat# 25030081 |
| Sodium pyruvate (100 mM) | Sigma-Aldrich | Cat# S8636 |
| D-Glucose | Sigma-Aldrich | Cat# G7021 |
| Oligomycin | Sigma-Aldrich | Cat# O4876 |
| FCCP | Sigma-Aldrich | Cat# C2920 |
| Antimycin A | Sigma-Aldrich | Cat# A8674 |
| Rotenone | Sigma-Aldrich | Cat# R8875 |
| 2-Deoxy-D-glucose | MilliporeSigma | Cat# 25972-1GM |
| Cu-ATSM | Sigma-Aldrich | Cat# SML0769 |
| LCS-1 (SOD1 inhibitor) | Sigma-Aldrich | Cat# 567417 |
| Sodium azide (NaN3) | Sigma-Aldrich | Cat# S2002 |
| DPBS | Corning | Cat# 21-031-CV |
| HEPES | Thermo Fisher Scientific | Cat# BP310-100 |
| D-Mannitol | Sigma-Aldrich | Cat# M4125 |
| Sucrose | Thermo Fisher Scientific | Cat# S5-3 |
| EDTA | Thermo Fisher Scientific | Cat# BP120-500 |
| EGTA | AMRESCO | Cat# 0732-100G |
| NP-40 | Sigma-Aldrich | Cat# I8896 |
| Triton X-100 | Sigma-Aldrich | Cat# T9284 |
| 4-(2-aminoethyl)benzenesulfonyl fluoride hydrochloride (AEBSF) | Sigma-Aldrich | Cat# 5.08436.0001 |
| Nitric Acid (Trace Metal Grade) | Fisher Chemical | Cat# A509P212 |
| UltraPure Distilled water | Invitrogen | Cat# 10977015 |
| Critical commercial assays | ||
| GeneArt CRISPR Nuclease Vector with OFP Reporter Kit | Thermo Fisher Scientific | Cat# A21174 |
| Cell Meter NIR Mitochondrial Membrane Potential Assay Kit | AAT Bioquest | Cat# 22802 |
| RNeasy Mini Kit | Qiagen | Cat# 74106 |
| ReadyScript® cDNA Synthesis Mix | Sigma-Aldrich | Cat# RDRT |
| PowerUp SYBR Green Master Mix | Thermo Fisher Scientific | Cat# A25741 |
| Seahorse XFe96/XF Pro FluxPak | Agilent | Cat# 103792-100 |
| Experimental models: Cell lines | ||
| Mouse embryonic fibroblasts (MEFs) | Wakabayashi et al.54 | N/A |
| Drp1-KO MEFs | Wakabayashi et al.54 | N/A |
| Slc25a3-KO MEFs | This paper | N/A |
| Drp1Slc25a3-KO MEFs | This paper | N/A |
| Oma1-KO MEFs | This paper | N/A |
| Drp1Oma1-KO MEFs | This paper | N/A |
| HEK293T cells | ATCC | CRL-3216 |
| Drp1-KO HEK293T cells | This paper | N/A |
| Oligonucleotides | ||
| esiRNA: control | Eupheria Biotech | Cat# EHURLUC-2.50UG |
| esiRNA: Tmem205 | Eupheria Biotech | Cat# EMU0105111UG |
| esiRNA: Tmem14c | Eupheria Biotech | Cat# EMU0178111UG |
| esiRNA: Tmem65 | Eupheria Biotech | Cat# EMU1989411UG |
| esiRNA: Slc25a3 | Eupheria Biotech | Cat# EMU0279111UG |
| esiRNA: Slc25a25 | Eupheria Biotech | Cat# EMU0492911UG |
| esiRNA: Slc25a24 | Eupheria Biotech | Cat# EMU0620011UG |
| esiRNA: Slc25a23 | Eupheria Biotech | Cat# EMU0811411UG |
| esiRNA: Slc25a41 | Eupheria Biotech | Cat# EMU0747211UG |
| esiRNA: Slc25a14 | Eupheria Biotech | Cat# EMU0148611UG |
| esiRNA: Romo1 | Eupheria Biotech | Cat# EMU0941811UG |
| esiRNA: Slc8b1 | Eupheria Biotech | Cat# EMU0563811UG |
| esiRNA: Mpv17 | Eupheria Biotech | Cat# EMU0414011UG |
| esiRNA: Ccdc51 | Eupheria Biotech | Cat# EMU0644611UG |
| esiRNA: Sfxn2 | Eupheria Biotech | Cat# EMU0852311UG |
| esiRNA: Sfxn4 | Eupheria Biotech | Cat# EMU0708411UG |
| esiRNA: Slc25a39 | Eupheria Biotech | Cat# EMU0863411UG |
| esiRNA: Slc25a27 | Eupheria Biotech | Cat# EMU0151411UG |
| esiRNA: Slc25a40 | Eupheria Biotech | Cat# EMU0715211UG |
| esiRNA: Ucp2 | Eupheria Biotech | Cat# EMU0245211UG |
| esiRNA: Slc25a43 | Eupheria Biotech | Cat# EMU0742911UG |
| esiRNA: Slc25a47 | Eupheria Biotech | Cat# EMU0050811UG |
| esiRNA: Mtch1 | Eupheria Biotech | Cat# EMU0129211UG |
| esiRNA: Mtch2 | Eupheria Biotech | Cat# EMU0506811UG |
| esiRNA: Mrs2 | Eupheria Biotech | Cat# EMU1814811UG |
| esiRNA: PXMP2 | Eupheria Biotech | Cat# EHU0368411UG |
| esiRNA: SLC25A30 | Eupheria Biotech | Cat# EHU0889811UG |
| esiRNA: SLC25A35 | Eupheria Biotech | Cat# EHU0158111UG |
| esiRNA: UCP1 | Eupheria Biotech | Cat# EHU0510211UG |
| esiRNA: UCP3 | Eupheria Biotech | Cat# EHU0730811UG |
| esiRNA: SLC25A34 | Eupheria Biotech | Cat# EHU1545611UG |
| shRNA targeting sequence: scramble: CCTAAGGTTAAGTCGCCCTCG | Addgene | 1864 |
| shRNA targeting sequence: Ccdc51 #1: GAAGAGAAGAGGCTCCGAATA | Sigma-Aldrich | TRCN0000251923 |
| shRNA targeting sequence: Ccdc51 #2: GTGAACAGGCTTCTAGCTATT | Sigma-Aldrich | TRCN0000248576 |
| shRNA targeting sequence: Mpv17 #1: CCCACGAATAGACACGCATTT | Sigma-Aldrich | TRCN0000120697 |
| shRNA targeting sequence: Mpv17 #2: GCTGGATCACTGATGGGCGTA | Sigma-Aldrich | TRCN0000120699 |
| shRNA targeting sequence: Slc25a3 #1: GCAACATACTTGGTGAGGAAA | Sigma-Aldrich | TRCN0000070020 |
| shRNA targeting sequence: Slc25a3 #2: CGACTCTGTGAAGGTCTACTT | Sigma-Aldrich | TRCN0000070021 |
| shRNA targeting sequence: Slc25a40: CATCCACCTCTAGATCATAAT | Sigma-Aldrich | TRCN0000432566 |
| CRISPR KO targeting sequence: human Drp1: GCCTGTAGGTGATCAACCTA | This paper | N/A |
| CRISPR KO targeting sequence: mouse Slc25a3: TTGGTGGGGTCTTAAGTTGT | This paper | N/A |
| CRISPR KO targeting sequence: mouse Oma1: GAGTGAATAACCTGGCCAAC | This paper | N/A |
| qPCR primers for Opa1-L1, Opa1-L2 and GAPDH | This paper | N/A |
| Recombinant DNA | ||
| pHR-CMV8.2ΔR | Addgene | 8455 |
| pHR-CMV8.9ΔR | Murata et al.19 | N/A |
| pCMV-VSVG | Addgene | 8454 |
| pHR-Sin Su9-GFP | Murata et al.19 | N/A |
| pLKO.1 scrambled shRNA | This paper | N/A |
| pLKO.1 Ccdc51 shRNA #1 | This paper | N/A |
| pLKO.1 Ccdc51 shRNA #2 | This paper | N/A |
| pLKO.1 Mpv17 shRNA #1 | This paper | N/A |
| pLKO.1 Mpv17 shRNA #2 | This paper | N/A |
| pLKO.1 Slc25a3 shRNA #1 | This paper | N/A |
| pLKO.1 Slc25a3 shRNA #2 | This paper | N/A |
| pLKO.1 Slc25a40 shRNA | This paper | N/A |
| pInducer20 human Opa1(L2)-HA | Murata et al.19 | N/A |
| mUNG1(Y147A) | Addgene | 70110 |
| pHR-Sin mUNG1(Y147A) | This paper | N/A |
| Human Slc25a3 | Addgene | 132040 |
| pHR-Sin mouse Slc25a3(WT) | This paper | N/A |
| pHR-Sin mouse Slc25a3(H75A) | This paper | N/A |
| pHR-Sin mouse Slc25a3(L175A) | This paper | N/A |
| pHR-Sin human Slc25a3(WT) | This paper | N/A |
| pHR-Sin human Slc25a3(G72E) | This paper | N/A |
| Human SOD1 | Wang et al.55 | N/A |
| pHR-Sin human SOD1 | This paper | N/A |
| Software and algorithms | ||
| MitoCarta3.0 | Rath et al.28 | www.broadinstitute.org/mitocarta |
| FIJI | NIH | https://imagej.net/software/fiji/downloads |
| Prism | GraphPad | https://www.graphpad.com/features |
| Design & Analysis Software | Thermo Fisher Scientific | https://www.thermofisher.com/us/en/home/technical-resources/software-downloads/quantstudio-3-5-real-time-pcr-systems.html |
| Syngistix for AA Software | PerkinElmer | https://www.perkinelmer.com/product/syngistix-for-aa-standard-sw-assembly-n1010302 |
Highlights.
Flickering prevents harmful extreme mitochondrial fusion by inducing Opa1 cleavage.
Mitochondrial copper, transported by Slc25a3, regulates flickering.
Copper-containing SOD1 suppresses flickering.
Copper-containing cytochrome c oxidase promotes flickering.
Acknowledgments
We are grateful to Dr. Alexander M. van der Bliek at UCLA and Dr. Jiou Wang at Johns Hopkins Bloomberg School of Public Health for providing the MFF antibody and the SOD1 plasmid, respectively. We also thank members of the Iijima and Sesaki labs for helpful discussions and technical assistance. This work was supported by NIH grants to MI (GM131768) and HS (GM144103).
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Declaration of interests
The authors declare no competing interests.
References
- 1.Murata D, Arai K, Iijima M, and Sesaki H (2020). Mitochondrial division, fusion and degradation. J Biochem 167, 233–241. 10.1093/jb/mvz106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Kameoka S, Adachi Y, Okamoto K, Iijima M, and Sesaki H (2018). Phosphatidic Acid and Cardiolipin Coordinate Mitochondrial Dynamics. Trends Cell Biol 28, 67–76. 10.1016/j.tcb.2017.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Kashatus DF (2018). The regulation of tumor cell physiology by mitochondrial dynamics. Biochem Biophys Res Commun 500, 9–16. 10.1016/j.bbrc.2017.06.192. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Friedman JR, and Nunnari J (2014). Mitochondrial form and function. Nature 505, 335–343. 10.1038/nature12985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Youle RJ, and van der Bliek AM (2012). Mitochondrial fission, fusion, and stress. Science 337, 1062–1065. 10.1126/science.1219855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Widlansky ME, and Hill RB (2018). Mitochondrial regulation of diabetic vascular disease: an emerging opportunity. Transl Res 202, 83–98. 10.1016/j.trsl.2018.07.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Roy M, Reddy PH, Iijima M, and Sesaki H (2015). Mitochondrial division and fusion in metabolism. Current opinion in cell biology 33C, 111–118. 10.1016/j.ceb.2015.02.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Itoh K, Nakamura K, Iijima M, and Sesaki H (2013). Mitochondrial dynamics in neurodegeneration. Trends in Cell Biology 23, 64–71. 10.1016/j.tcb.2012.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Liesa M, and Shirihai OS (2013). Mitochondrial dynamics in the regulation of nutrient utilization and energy expenditure. Cell metabolism 17, 491–506. 10.1016/j.cmet.2013.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Serasinghe MN, and Chipuk JE (2017). Mitochondrial Fission in Human Diseases. Handb Exp Pharmacol. 10.1007/164_2016_38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Mishra P, and Chan DC (2014). Mitochondrial dynamics and inheritance during cell division, development and disease. Nature reviews. Molecular cell biology 15, 634–646. 10.1038/nrm3877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Tamura Y, Itoh K, and Sesaki H (2011). SnapShot: Mitochondrial dynamics. Cell. 145, 1158. 10.1016/j.cell.2011.06.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Kraus F, and Ryan MT (2017). The constriction and scission machineries involved in mitochondrial fission. J Cell Sci 130, 2953–2960. 10.1242/jcs.199562. [DOI] [PubMed] [Google Scholar]
- 14.Pernas L, and Scorrano L (2016). Mito-Morphosis: Mitochondrial Fusion, Fission, and Cristae Remodeling as Key Mediators of Cellular Function. Annual review of physiology 78, 505–531. 10.1146/annurev-physiol-021115-105011. [DOI] [PubMed] [Google Scholar]
- 15.Prudent J, and McBride HM (2017). The mitochondria-endoplasmic reticulum contact sites: a signalling platform for cell death. Curr Opin Cell Biol 47, 52–63. 10.1016/j.ceb.2017.03.007. [DOI] [PubMed] [Google Scholar]
- 16.van der Bliek AM, Shen Q, and Kawajiri S (2013). Mechanisms of mitochondrial fission and fusion. Cold Spring Harb Perspect Biol 5. 10.1101/cshperspect.a011072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Ramachandran R (2018). Mitochondrial dynamics: The dynamin superfamily and execution by collusion. Semin Cell Dev Biol 76, 201–212. 10.1016/j.semcdb.2017.07.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.McNiven MA, Cao H, Pitts KR, and Yoon Y (2000). The dynamin family of mechanoenzymes: pinching in new places. Trends Biochem Sci 25, 115–120. [DOI] [PubMed] [Google Scholar]
- 19.Murata D, Yamada T, Tokuyama T, Arai K, Quiros PM, Lopez-Otin C, Iijima M, and Sesaki H (2020). Mitochondrial Safeguard: a stress response that offsets extreme fusion and protects respiratory function via flickering-induced Oma1 activation. EMBO J 39, e105074. 10.15252/embj.2020105074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Yamada T, Murata D, Adachi Y, Itoh K, Kameoka S, Igarashi A, Kato T, Araki Y, Huganir RL, Dawson TM, et al. (2018). Mitochondrial Stasis Reveals p62-Mediated Ubiquitination in Parkin-Independent Mitophagy and Mitigates Nonalcoholic Fatty Liver Disease. Cell Metab 28, 588–604. 10.1016/j.cmet.2018.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Yamada T, Dawson TM, Yanagawa T, Iijima M, and Sesaki H (2019). SQSTM1/p62 promotes mitochondrial ubiquitination independently of PINK1 and PRKN/parkin in mitophagy. Autophagy 15, 2012–2018. 10.1080/15548627.2019.1643185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Tabara LC, and Prudent J (2020). The last wall of defense to prevent extreme and deleterious mitochondrial fusion. EMBO J 39, e107326. 10.15252/embj.2020107326. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Chalmers S, Saunter CD, Girkin JM, and McCarron JG (2015). Flicker-assisted localization microscopy reveals altered mitochondrial architecture in hypertension. Sci Rep 5, 16875. 10.1038/srep16875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Duchen MR, Leyssens A, and Crompton M (1998). Transient mitochondrial depolarizations reflect focal sarcoplasmic reticular calcium release in single rat cardiomyocytes. J Cell Biol 142, 975–988. 10.1083/jcb.142.4.975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Loew LM, Tuft RA, Carrington W, and Fay FS (1993). Imaging in five dimensions: time-dependent membrane potentials in individual mitochondria. Biophys J 65, 2396–2407. 10.1016/S0006-3495(93)81318-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Roy M, Itoh K, Iijima M, and Sesaki H (2016). Parkin suppresses Drp1-independent mitochondrial division. Biochem Biophys Res Commun 475, 283–288. 10.1016/j.bbrc.2016.05.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Lee H, and Yoon Y (2014). Transient contraction of mitochondria induces depolarization through the inner membrane dynamin OPA1 protein. The Journal of biological chemistry 289, 11862–11872. 10.1074/jbc.M113.533299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Rath S, Sharma R, Gupta R, Ast T, Chan C, Durham TJ, Goodman RP, Grabarek Z, Haas ME, Hung WHW, et al. (2021). MitoCarta3.0: an updated mitochondrial proteome now with sub-organelle localization and pathway annotations. Nucleic Acids Res 49, D1541–D1547. 10.1093/nar/gkaa1011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Yang D, Buchholz F, Huang Z, Goga A, Chen CY, Brodsky FM, and Bishop JM (2002). Short RNA duplexes produced by hydrolysis with Escherichia coli RNase III mediate effective RNA interference in mammalian cells. Proc Natl Acad Sci U S A 99, 9942–9947. 10.1073/pnas.152327299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Cobine PA, Moore SA, and Leary SC (2021). Getting out what you put in: Copper in mitochondria and its impacts on human disease. Biochim Biophys Acta Mol Cell Res 1868, 118867. 10.1016/j.bbamcr.2020.118867. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Bhoj EJ, Li M, Ahrens-Nicklas R, Pyle LC, Wang J, Zhang VW, Clarke C, Wong LJ, Sondheimer N, Ficicioglu C, and Yudkoff M (2015). Pathologic Variants of the Mitochondrial Phosphate Carrier SLC25A3: Two New Patients and Expansion of the Cardiomyopathy/Skeletal Myopathy Phenotype With and Without Lactic Acidosis. JIMD Rep 19, 59–66. 10.1007/8904_2014_364. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Boulet A, Vest KE, Maynard MK, Gammon MG, Russell AC, Mathews AT, Cole SE, Zhu X, Phillips CB, Kwong JQ, et al. (2018). The mammalian phosphate carrier SLC25A3 is a mitochondrial copper transporter required for cytochrome c oxidase biogenesis. J Biol Chem 293, 1887–1896. 10.1074/jbc.RA117.000265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Zhu X, Boulet A, Buckley KM, Phillips CB, Gammon MG, Oldfather LE, Moore SA, Leary SC, and Cobine PA (2021). Mitochondrial copper and phosphate transporter specificity was defined early in the evolution of eukaryotes. Elife 10. 10.7554/eLife.64690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.MacVicar T, and Langer T (2016). OPA1 processing in cell death and disease - the long and short of it. J Cell Sci 129, 2297–2306. 10.1242/jcs.159186. [DOI] [PubMed] [Google Scholar]
- 35.Head B, Griparic L, Amiri M, Gandre-Babbe S, and van der Bliek AM (2009). Inducible proteolytic inactivation of OPA1 mediated by the OMA1 protease in mammalian cells. The Journal of cell biology 187, 959–966. 10.1083/jcb.200906083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Ehses S, Raschke I, Mancuso G, Bernacchia A, Geimer S, Tondera D, Martinou JC, Westermann B, Rugarli EI, and Langer T (2009). Regulation of OPA1 processing and mitochondrial fusion by m-AAA protease isoenzymes and OMA1. The Journal of cell biology 187, 1023–1036. 10.1083/jcb.200906084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Ishihara T, Ban-Ishihara R, Maeda M, Matsunaga Y, Ichimura A, Kyogoku S, Aoki H, Katada S, Nakada K, Nomura M, et al. (2015). Dynamics of mitochondrial DNA nucleoids regulated by mitochondrial fission is essential for maintenance of homogeneously active mitochondria during neonatal heart development. Mol Cell Biol 35, 211–223. 10.1128/MCB.01054-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Ban-Ishihara R, Ishihara T, Sasaki N, Mihara K, and Ishihara N (2013). Dynamics of nucleoid structure regulated by mitochondrial fission contributes to cristae reformation and release of cytochrome c. Proceedings of the National Academy of Sciences of the United States of America 110, 11863–11868. 10.1073/pnas.1301951110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Itoh K, Tamura Y, Iijima M, and Sesaki H (2013). Effects of Fcj1-Mos1 and mitochondrial division on aggregation of mitochondrial DNA nucleoids and organelle morphology. Molecular biology of the cell 24, 1842–1851. 10.1091/mbc.E13-03-0125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Lewis SC, Uchiyama LF, and Nunnari J (2016). ER-mitochondria contacts couple mtDNA synthesis with mitochondrial division in human cells. Science 353, aaf5549. 10.1126/science.aaf5549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Kageyama Y, Hoshijima M, Seo K, Bedja D, Sysa-Shah P, Andrabi SA, Chen W, Hoke A, Dawson VL, Dawson TM, et al. (2014). Parkin-independent mitophagy requires Drp1 and maintains the integrity of mammalian heart and brain. The EMBO journal 33, 2798–2813. 10.15252/embj.201488658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Spadafora D, Kozhukhar N, Chouljenko VN, Kousoulas KG, and Alexeyev MF (2016). Methods for Efficient Elimination of Mitochondrial DNA from Cultured Cells. PLoS One 11, e0154684. 10.1371/journal.pone.0154684. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Ruiz LM, Libedinsky A, and Elorza AA (2021). Role of Copper on Mitochondrial Function and Metabolism. Front Mol Biosci 8, 711227. 10.3389/fmolb.2021.711227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Mayr JA, Zimmermann FA, Horvath R, Schneider HC, Schoser B, Holinski-Feder E, Czermin B, Freisinger P, and Sperl W (2011). Deficiency of the mitochondrial phosphate carrier presenting as myopathy and cardiomyopathy in a family with three affected children. Neuromuscul Disord 21, 803–808. 10.1016/j.nmd.2011.06.005. [DOI] [PubMed] [Google Scholar]
- 45.Mayr JA, Merkel O, Kohlwein SD, Gebhardt BR, Bohles H, Fotschl U, Koch J, Jaksch M, Lochmuller H, Horvath R, et al. (2007). Mitochondrial phosphate-carrier deficiency: a novel disorder of oxidative phosphorylation. Am J Hum Genet 80, 478–484. 10.1086/511788. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Seifert EL, Gal A, Acoba MG, Li Q, Anderson-Pullinger L, Golenar T, Moffat C, Sondheimer N, Claypool SM, and Hajnoczky G (2016). Natural and Induced Mitochondrial Phosphate Carrier Loss: DIFFERENTIAL DEPENDENCE OF MITOCHONDRIAL METABOLISM AND DYNAMICS AND CELL SURVIVAL ON THE EXTENT OF DEPLETION. J Biol Chem 291, 26126–26137. 10.1074/jbc.M116.744714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Kaplan JH, and Maryon EB (2016). How Mammalian Cells Acquire Copper: An Essential but Potentially Toxic Metal. Biophys J 110, 7–13. 10.1016/j.bpj.2015.11.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Chen L, Min J, and Wang F (2022). Copper homeostasis and cuproptosis in health and disease. Signal Transduct Target Ther 7, 378. 10.1038/s41392-022-01229-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Halliwell B (2024). Understanding mechanisms of antioxidant action in health and disease. Nat Rev Mol Cell Biol 25, 13–33. 10.1038/s41580-023-00645-4. [DOI] [PubMed] [Google Scholar]
- 50.Sies H, and Jones DP (2020). Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat Rev Mol Cell Biol 21, 363–383. 10.1038/s41580-020-0230-3. [DOI] [PubMed] [Google Scholar]
- 51.Murphy MP (2009). How mitochondria produce reactive oxygen species. Biochem J 417, 1–13. 10.1042/BJ20081386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Kageyama Y, Zhang Z, Roda R, Fukaya M, Wakabayashi J, Wakabayashi N, Kensler TW, Reddy PH, Iijima M, and Sesaki H (2012). Mitochondrial division ensures the survival of postmitotic neurons by suppressing oxidative damage. J Cell Biol 197, 535–551. 10.1083/jcb.201110034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Gandre-Babbe S, and van der Bliek AM (2008). The novel tail-anchored membrane protein Mff controls mitochondrial and peroxisomal fission in mammalian cells. Mol Biol Cell 19, 2402–2412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Wakabayashi J, Zhang Z, Wakabayashi N, Tamura Y, Fukaya M, Kensler TW, Iijima M, and Sesaki H (2009). The dynamin-related GTPase Drp1 is required for embryonic and brain development in mice. J Cell Biol 186, 805–816. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Wang J, Farr GW, Zeiss CJ, Rodriguez-Gil DJ, Wilson JH, Furtak K, Rutkowski DT, Kaufman RJ, Ruse CI, Yates JR 3rd, et al. (2009). Progressive aggregation despite chaperone associations of a mutant SOD1-YFP in transgenic mice that develop ALS. Proc Natl Acad Sci U S A 106, 1392–1397. 10.1073/pnas.0813045106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Senoo H, Murata D, Wai M, Arai K, Iwata W, Sesaki H, and Iijima M (2021). KARATE: PKA-induced KRAS4B-RHOA-mTORC2 supercomplex phosphorylates AKT in insulin signaling and glucose homeostasis. Mol Cell 81, 4622–4634 e4628. 10.1016/j.molcel.2021.09.001. [DOI] [PubMed] [Google Scholar]
- 57.Yeung N, Murata D, Iijima M, and Sesaki H (2023). Role of human HSPE1 for OPA1 processing independent of HSPD1. iScience 26, 106067. 10.1016/j.isci.2023.106067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Adachi Y, Itoh K, Yamada T, Cerveny KL, Suzuki TL, Macdonald P, Frohman MA, Ramachandran R, Iijima M, and Sesaki H (2016). Coincident Phosphatidic Acid Interaction Restrains Drp1 in Mitochondrial Division. Mol Cell 63, 1034–1043. 10.1016/j.molcel.2016.08.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Yamada T, Adachi Y, Fukaya M, Iijima M, and Sesaki H (2016). Dynamin-Related Protein 1 Deficiency Leads to Receptor-Interacting Protein Kinase 3-Mediated Necroptotic Neurodegeneration. Am J Pathol 86, 2798–2802. 10.1016/j.ajpath.2016.06.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All data reported in this paper will be shared by the lead contact upon request.
This paper does not report original code.
Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request.
