Abstract
Spontaneous Ca2+ signaling from the InsP3R intracellular Ca2+ release channel to mitochondria is essential for optimal oxidative phosphorylation (OXPHOS) and ATP production. In cells with defective OXPHOS, reductive carboxylation replaces oxidative metabolism to maintain amounts of reducing equivalents and metabolic precursors. To investigate the role of mitochondrial Ca2+ uptake in regulating bioenergetics in these cells, we used OXPHOS-competent and OXPHOS-defective cells. Inhibition of InsP3R activity or mitochondrial Ca2+ uptake increased α-ketoglutarate (αKG) abundance and the NAD+/NADH ratio, indicating that constitutive endoplasmic reticulum (ER)–to–mitochondria Ca2+ transfer promoted optimal αKG dehydrogenase (αKGDH) activity. Reducing mitochondrial Ca2+ inhibited αKGDH activity and increased NAD+, which induced SIRT1-dependent autophagy in both OXPHOS-competent and OXPHOS-defective cells. Whereas autophagic flux in OXPHOS-competent cells promoted cell survival, it was impaired in OXPHOS-defective cells because of inhibition of autophagosome-lysosome fusion. Inhibition of αKGDH and impaired autophagic flux in OXPHOS-defective cells resulted in pronounced cell death in response to interruption of constitutive flux of Ca2+ from ER to mitochondria. These results demonstrate that mitochondria play a fundamental role in maintaining bioenergetic homeostasis of both OXPHOS-competent and OXPHOS-defective cells, with Ca2+ regulation of αKGDH activity playing a pivotal role. Inhibition of ER-to-mitochondria Ca2+ transfer may represent a general therapeutic strategy against cancer cells regardless of their OXPHOS status.
INTRODUCTION
Many cancers switch from oxidative phosphorylation (OXPHOS) to glycolysis as the main source of adenosine triphosphate (ATP) generation in the presence of ample oxygen, a phenomenon known as the Warburg effect (1). Although enhanced aerobic glycolysis has been misinterpreted as a sign of mitochondrial dysfunction in cancer cells, fully functional mitochondria are needed in many cancer cell types (2–5). Oncogene activation increases mitochondrial metabolism, and anchorage-independent growth and metastatic potential are correlated with mitochondrial function (2, 6). In addition to its role in ATP production by OXPHOS, mitochondria fuel biosynthetic pathways through the tricarboxylic acid (TCA) cycle to produce lipids, proteins, and nucleic acids required in highly proliferating cells (3, 4, 7–11). Many studies have targeted OXPHOS as a therapeutic strategy in cancer (12–15).
In rapidly proliferating cells and many cancer cells as well as in cells with defective OXPHOS, including those caused by mutations in either fumarate hydratase or components of succinate dehydrogenase, reductive carboxylation of glutamine-derived α-ketoglutarate (αKG) by nicotinamide adenine dinucleotide (NAD+) phosphate (NADP+)/reduced form of NADP+ (NADPH)–dependent isocitrate dehydrogenase (IDH1 and IDH2) generates citrate. Citrate, in turn, provides acetyl–coenzyme A (CoA) for lipid synthesis and for replenishing TCA cycle intermediates (16, 17), including aspartate for nucleotide synthesis, by its cleavage to form oxaloacetate and its transamination (18, 19). In OXPHOS-defective 143BΔcytb cybrid cells and other cells with defective OXPHOS, the reducing equivalents needed for the reductive carboxylation reaction are generated by oxidation of αKG by αKG dehydrogenase (αKGDH). Thus, oxidative metabolism persists and coexists in 143BΔcytb cells and other cell lines that use reductive carboxylation (17, 20). αKGDH activity is under complex regulation by substrate levels and cofactors (21), including mitochondrial matrix Ca2+ concentration. Ca2+ increases the affinity of αKGDH for its substrate, enhancing its activity over 100-fold (22).
Low-level constitutive Ca2+ release from the endoplasmic reticulum (ER) through inositol 1,4,5-trisphosphate receptors (InsP3Rs) is essential to maintain OXPHOS (23). Inhibition of InsP3R activity reduces the O2 consumption rate (OCR), lowers cell ATP levels, and activates adenosine monophosphate (AMP)–activated protein kinase (AMPK)–dependent, mechanistic target of rapamycin (mTOR)–independent prosurvival autophagy. A similar and nonadditive response is elicited by inhibition of mitochondrial Ca2+ uptake by the Ca2+ uniporter (24–26), the major pathway for Ca2+ uptake into the mitochondrial matrix (27). Thus, constitutive low-level ER-to-mitochondrial Ca2+ transfer is required to preserve normal bioenergetics by regulation of OXPHOS. Here, we demonstrated that constitutive InsP3R-mediated Ca2+ signaling to mitochondria was required to maintain mitochondrial bioenergetics even in the absence of OXPHOS. In cells with defective OXPHOS, constitutive low-level mitochondrial Ca2+ uptake sustained the oxidative activity of the αKGDH complex to maintain reduced form of NAD+ (NADH) levels to support reductive carboxylation. In its absence, NAD+/NADH levels increased and autophagy was activated in a SIRT1-dependent manner. Autophagy promoted survival in cells with competent OXPHOS, but it was insufficient in OXPHOS-defective cells because of defective autophagosome-lysosome fusion. These results indicate that constitutive ER-to-mitochondria Ca2+ transfer regulates cellular bioenergetics regardless of the OXPHOS status and suggest that targeting mitochondrial Ca2+ uptake may provide a strategy to kill cancer cells by impinging on multiple pathways regulating mitochondrial metabolism.
RESULTS
Viability of OXPHOS-defective cells requires constitutive InsP3R-mediated Ca2+ release
To determine the role of constitutive low-level InsP3R-mediated Ca2+ release on bioenergetics of cells with defective OXPHOS, we used cybrids derived from 143B osteosarcoma cells in which mitochondrial DNA had been depleted (p0) and then reconstituted with either wild-type mitochondrial DNA (143Bwt) or with mitochondrial DNA containing a 4–base pair mutation in the cytochrome b gene (143BΔcytb). In the 143BΔcytb cells, the flow of electrons stops at complex III (6), rendering mitochondria unable to reduce O2 and produce ATP. Compared with 143Bwt cells, 143BΔcytb cells had lower basal OCR (Fig. 1A) but higher basal extracellular acidification rate (ECAR), a measure of glycolysis (Fig. 1B), as expected. Although 143Bwt cells are cancer derived and activation of the Warburg effect was expected, they nevertheless use mitochondria for OXPHOS, as observed in other cancer cells (8, 9, 12, 24, 28–31). ATP content in basal conditions was similar in 143Bwt and 143BΔcytb cells (Fig. 1C). Inhibition of respiratory complexes III or V with antimycin A and oligomycin, respectively, did not affect ATP content in 143BΔcytb cells, whereas 2-deoxyglucose (2DG) did, confirming that glycolysis provided the main source of ATP in these cells. In contrast, these inhibitors reduced ATP levels in 143Bwt cells (Fig. 1C).
Fig. 1. OXPHOS-defective 143BΔcytb cells rely on constitutive InsP3R-mediated Ca2+ signals.
(A) Basal oxygen consumption rate (OCR) of cells incubated with 5 μM XeB for 4 hours. (B) Extracellular acidification rate (ECAR) of cells incubated with 5 μM XeB for 4 hours. Data in (A) and (B) represent means ± SEM of three independent experiments with 10 replicates each. **P < 0.01 and ***P < 0.001 compared to respective control. NS, not significant. (C) Cells were incubated with 5 μM XeB, 1 μM antimycin A (AA) or 1 μM oligomycin (Oli), 2-deoxyglucose (2DG; 20 mM), or vehicle control (C) for 4 hours, and intracellular ATP levels were determined. Means ± SEM of three independent experiments performed in triplicate. ***P < 0.001 compared to control. (D) Representative recordings of Ca2+ release events in unstimulated cells treated or not with 5 μM XeB for 30 min. (E) Quantification of Ca2+ release events. Means ± SEM of three independent experiments. From each experiment, 30 cells were analyzed. (F) Representative traces of [Ca2+]c (green) and [Ca2+]m (red) responses in 143Bwt and 143BΔcytb cells challenged with histamine (100 μM). (G) Quantification of peak rhod-2 fluorescence. Means ± SEM of three independent experiments. In each experiment, 100 cells were analyzed. (H) Cells were treated with 2.5 or 5 μM XeB for 12 hours, and death was determined by LDH release. Means ± SEM of three independent experiments each with six replicates. **P < 0.01 and ***P < 0.001.
Like normal cells (23), cancer cells require constitutive InsP3R-mediated Ca2+ transfer to mitochondria for optimal mitochondrial function and viability (24). Both 143Bwt and 143BΔcytb cells exhibited similar constitutive Ca2+ release activity that was InsP3R dependent, as shown by the suppression of this activity by the specific InsP3R inhibitor xestospongin B (XeB) (Fig. 1, D and E) (32). The cytoplasmic free Ca2+ concentration ([Ca2+]i) and mitochondrial matrix free Ca2+ concentration ([Ca2+]m) responses to histamine stimulation were similar in 143BΔcytb and 143Bwt cells (Fig. 1, F and G) despite a lower mitochondrial membrane potential (Δψm) in the 143BΔcytb cells (fig. S1A), as has been previously observed (33). Knockdown of mitochondrial calcium uniporter (MCU), the pore-forming subunit of the mitochondrial Ca2+ uniporter, prevented mitochondrial Ca2+ uptake (fig. S1, B and C). Labeling with nonyl acridine orange (NAO) revealed no differences in the numbers of mitochondria betweeen143Bwt and 143BΔcytb cells (fig. S1D). XeB decreased the OCR and ATP levels in 143Bwt cells, as observed in other cancer and normal cell types (23, 24) but not in 143BΔcytb cells (Fig. 1, A and C). XeB did not change the rate of ECAR in either cell line (Fig. 1B). In tumorigenic breast and prostate cells with functional mitochondria, prolonged inhibition of InsP3R by XeB generates a bioenergetic crisis that results in >70% cell death (24). Because 143BΔcytb cells rely mainly on glycolysis, it was expected that they would be resistant to XeB-induced death. However, XeB caused cytotoxicity, with the 143BΔcytb cells significantly more sensitive than 143Bwt cells (Fig. 1H).
To confirm that the observed effects of XeB were specifically due to inhibition of InsP3R activity, types 1 and 3 InsP3Rs were simultaneously knocked down by mixing pools of four small interfering RNAs (siRNAs) against each of the types 1 and 3 InsP3Rs (IP3R1 + 3), generating >80 and 90% knockdown in 143Bwt and 143BΔcytb cells, respectively (Fig. 2A). InsP3R knockdown reduced basal OCR in 143Bwt cells but not in 143BΔcytb cells (Fig. 2B). In contrast, cell death was enhanced in both lines, with 143BΔcytb cells being more sensitive (Fig. 2C), similar to their greater sensitivity to XeB. Similar results were observed when types 1 and 3 InsP3Rs were simultaneously knocked down with a single siRNA designed to recognize both isoforms (IP3R1 and IP3R3) (fig. S1, E to G). These results suggest that 143BΔcytb cells, despite defective OXPHOS, require low-level constitutive InsP3R Ca2+ release for survival. The type 3 InsP3R has been associated with apoptosis resistance in cancer (34, 35). However, knockdown of only this isoform (>80% in both cell lines; fig. S1, H and I) did not affect basal OCR (fig. S1J) or cell viability (fig. S1K) in either cell line, in agreement with the previous observation that the expression of either type 1 or type 3 InsP3R is sufficient to maintain cellular bioenergetics (23). To determine if InsP3R activity provided Ca2+ for mitochondrial uptake, we knocked down (~80%) the expression of MCU with two different siRNAs (Fig. 2, D and E). Knockdown of MCU with each siRNA reduced OCR in 143Bwt cells but not in 143BΔcytb cells (Fig. 2, F and G), phenocopying the effects of XeB and InsP3R knockdown. In both cell lines, significant cell death was observed, which was again more profound in the 143BΔcytb cells (Fig. 2, H and I), exceeding that caused by InsP3R inhibition. These results suggest that although 143BΔcytb cells do not use mitochondria to generate ATP, InsP3R-mediated Ca2+ transfer to mitochondria is nevertheless essential to maintain their viability.
Fig. 2. Simultaneous knockdown of types 1 and 3 InsP3R or mitochondrial Ca2+ uptake induces cell death in OXPHOS-defective 143BΔcytb cells.
(A) Representative Western blots of types 1 and 3 InsP3R in cells transiently transfected with a pool of four siRNAs against type 1 InsP3R and a pool of four siRNAs against type 3 InsP3R (IP3R1 + 3) or a nontargeting (NT) siRNA for 24 hours. (B) Basal OCR of cells transiently transfected with siRNAs against types 1 and 3 InsP3R (IP3R1 + 3) or nontargeting siRNA for 24 hours. Means ± SEM of three independent experiments with 10 replicates each. ***P < 0.001 compared to control. (C) Cell death as assessed by LDH release in cells transiently transfected with siRNAs targeting types 1 and 3 InsP3R (IP3R1 + 3) or nontargeting siRNA for 24 hours. Means ± SEM of three independent experiment each with six replicates. ***P < 0.001. (D) Representative Western blots of MCU in cells transiently transfected with MCU siRNA#1 or nontargeting siRNA for 72 hours. (E) Representative Western blots of MCU in cells transiently transfected with MCU siRNA#2 or nontargeting siRNA for 72 hours. (F) Basal OCR of cells transiently transfected with MCU siRNA#1 or nontargeting siRNA for 72 hours. (G) Basal OCR of cells transiently transfected with MCU siRNA#2 or nontargeting siRNA for 72 hours. (H) Cell death as assessed by LDH release in cells transiently transfected with MCU siRNA#1 or a nontargeting siRNA for 72 hours. (I) Cell death as assessed by LDH release in cells transiently transfected with MCU siRNA#2 or a nontargeting siRNA for 72 hours. In (A), bar graphs represent IP3R1 and IP3R3 expression normalized to tubulin. In (D) and (E), bar graphs represent MCU expression normalized to tubulin. Data in (A), (D), and (E) are means ± SEM of three independent biological replicates. **P < 0.01 and ***P < 0.001 compared with respective controls (NT). Data in (B), (F), and (G) represent means ± SEM of three independent experiments with 10 replicates each. **P < 0.01 and ***P < 0.001 compared to respective controls (NT). Cytotoxicity data in (C), (H), and (I) are means ± SEM of three independent experiments, each with six replicates. **P < 0.01 and ***P < 0.001 compared with respective controls (NT).
XeB reduced Δψm in 143Bwt cells but not in 143BΔcytb cells (fig. S1A), which suggested that XeB-induced cytotoxicity of 143BΔcytb cells was not caused by loss of Δψm. 143BΔcytb cells have been reported to have higher resting levels of reactive oxygen species (ROS) compared with 143Bwt cells (36). Although increased ROS generation can cause cell death (37), we detected no differences in ROS levels between143Bwt and 143BΔcytb cells in resting and XeB-treated conditions (fig. S1L), in agreement with other reports (6, 38). Furthermore, XeB-induced cell death was not affected by the ROS scavenger N-acetylcysteine (NAC) (fig. S1M).
Although 143BΔcytb cells have defective OXPHOS, they require the TCA cycle for glutamine metabolism through reductive carboxylation (6). We therefore asked whether the requirement for constitutive Ca2+ transfer from the ER to mitochondria was specific to the 143BΔcytb cells or if it could be recapitulated in other cells that exhibit OXPHOS deficiency. The RCC4 human renal carcinoma cell line has a von Hippel–Lindau (vhl) mutation that results in constitutive activation of hypoxia-inducible factors, repression of mitochondrial function and OCR, and activation of reductive carboxylation (fig. S2A) (39). RCC4 cells were significantly more sensitive to XeB-induced toxicity than RCC4 cells expressing normal vhl, with cell death reaching ~70% after 12 hours (fig. S2, B and C). XeB-sensitive basal Ca2+ release activity was similar in RCC4 cells and RCC4 cells expressing vhl (fig. S2D). Thus, constitutive Ca2+ transfer from ER to mitochondria is required in both 143BΔcytb and RCC4 cells that exhibit OXPHOS deficiency.
Inhibition of constitutive InsP3R-mediated Ca2+ release suppresses αKGDH activity
Like many cancer cells, 143BΔcytb cells rely on reductive carboxylation to generate metabolites for growth and proliferation (16–19). When oxidation of metabolites in the TCA cycle is unfavorable, such as when OXPHOS activity is defective, glutamine provides glutamate through glutaminase activity, which is converted by glutamate dehydrogenases or aminotransferases to αKG (2). αKG is oxidized by αKGDH to succinyl-CoA to generate the reducing equivalents for IDH2-mediated reductive carboxylation of αKG to isocitrate (20) and other intermediates for production of TCA cycle metabolites (40). In addition, the reductive carboxylation of glutamine fuels compensatory aspartate synthesis from αKG through a pathway that requires the aspartate aminotransferase GOT1 (18, 19). Removal of glutamine was cytotoxic for 143BΔcytb cells, which was rescued by the addition of dimethyl-α-ketoglutarate (Fig. 3A), which can be readily used by αKGDH to sustain levels of metabolic intermediates. Furthermore, preventing the conversion of glutamate to αKG by inhibiting transaminases or glutamate dehydrogenase 1 with aminooxyacetate (AOA) or epigallocatechin gallate (EGCG), respectively, was much more cytotoxic for 143BΔcytb cells than for 143Bwt cells (Fig. 3, B and C). Unexpectedly, inhibition of glutaminase with CB839 reduced the viability of both cell lines (Fig. 3D). Nevertheless, these results indicate a reliance of 143BΔcytb cells on αKGDH activity, likely for the generation of NADH to maintain reductive carboxylation. In agreement, inhibition of αKGDH activity with CPI-613 had a significantly higher cytotoxic effect on 143BΔcytb cells (Fig. 3E).
Fig. 3. Inhibition of constitutive InsP3R-mediated Ca2+ release reduces αKGDH activity, enhances the NAD+/NADH ratio, and increases cell death that can be rescued by metabolic intermediates.
(A) Cells were cultured in normal media (control), glutamine-free media (−Gln), or glutamine-free media supplemented with 5 mM dimethyl-α-ketoglutarate (−Gln + αKG), and cell death was determined by propidium iodide incorporation. Means ± SEM of three independent experiments, each in triplicate. ***P < 0.001 compared to respective control. (B) Cells were treated with 1 mM aminotransferase inhibitor aminooxyacetate (AOA), and cell death was determined by propidium iodide incorporation. Means ± SEM of three independent experiments, each in triplicate. **P < 0.01 compared to respective control. (C) Cells were treated with 50 μM glutamate dehydrogenase (GDH) inhibitor epigallocatechin gallate (EGCG), and cell death was determined by propidium iodide incorporation. Means ± SEM of three independent experiments, each in triplicate. **P < 0.01 compared to respective control. (D) Cells were treated with 5 μM glutaminase inhibitor CB839, and cell death was determined by propidium iodide incorporation. Means ± SEM of three independent experiments, each in triplicate. **P < 0.01 compared to respective control. (E) Cells were treated with 0.5 mM αKGDH inhibitor CPI-613, and cell death was determined by propidium iodide incorporation. Means ± SEM of three independent experiments, each in triplicate. *P < 0.05 compared to respective control. (F) Cells were treated with 5 μM XeB (4 hours), and αKG levels were determined. Means ± SEM of three independent experiments each performed in triplicate. **P < 0.01 and ***P < 0.001 compared to respective control. (G) 143B cells transfected with a WT E1 subunit of the αKGDH or the D154A or E179A mutants were cultured with glutamine as the only carbon source for 24 hours, and cell death was determined by propidium iodide incorporation. Means ± SEM of three independent experiments, each in triplicate. ***P < 0.001 compared to respective control. (H) Cells were treated with 5 μM XeB for 1 and 4 hours or with 1 μM FCCP for 1 hour, and the NAD+/NADH ratio was determined. Means ± SEM of three independent experiments each performed in triplicate. ***P < 0.001 compared to respective control. (I) Representative maps of redox ratio in 143Bwt and 143BΔcytb treated or not with 5 μM XeB for 4 hours. (J) Bar graph represents quantification of the redox ratio. Means ± SD of three independent experiments. *P < 0.05 and **P < 0.01 compared to respective control. (K) Cell death assessed by LDH release in cells treated with 5 μM XeB for 12 hours in the presence of 5 mM dimethyl-α-ketoglutarate (αKG) or 5 mM methyl-pyruvate (Pyr). Means ± SEM of three independent experiments each performed in triplicate. ***P < 0.001 compared to respective control.
Because αKGDH is Ca2+ sensitive, we hypothesized that the cytotoxic effects of blocking ER-to-mitochondrial Ca2+ transfer in cells reliant on reductive carboxylation was caused by suppression of αKGDH activity. Exposure to XeB resulted in accumulation of αKG (Fig. 3F), consistent with a requirement of αKGDH activity on constitutive mitochondrial Ca2+ uptake, as observed for pyruvate dehydrogenase (PDH) (23). This result suggested that the inhibition of ER Ca2+ transfer to the mitochondria likely induced metabolic remodeling. A previous metabolomics study comparing the abundance of more than 90 metabolites between 143Bwt and 143BΔcytb revealed relatively minor differences between the two cell lines (20). We performed a metabolomics study to compare the abundance of metabolites from glycolysis, the pentose phosphate pathway, one-carbon/nucleotide metabolism, the TCA cycle, and other pathways in 143BΔcytb cells treated or not with XeB for 4 hours. Principal components analysis (fig. S3A) and hierarchical clustering (fig. S3B) revealed significant differences between the two groups. Analyses of metabolic intermediates generated by reductive carboxylation downstream of αKGDH revealed that succinate, malic acid, and citrate-isocitrate were significantly reduced after XeB treatment (Table 1), consistent with a decrease in αKGDH activity. These results suggest that inhibition of ER to mitochondria Ca2+ transfer alters the metabolic profile of 143BΔcytb cells, including a reduction in αKGDH activity.
Table 1. Inhibition of constitutive InsP3R-mediated Ca2+ release decreases reductive carboxylation metabolites downstream of αKGDH.
Untargeted metabolomics was performed on OXPHOS-defective 143BΔcytb cells treated or not with 5 μM XeB for 4 hours, and the fold change of relevant reductive carboxylation metabolites related to αKGDH were analyzed. Means ± SEM of three independent experiments each with three replicates.
Metabolite | Fold change | P value |
---|---|---|
Malic acid | 0.70 ± 0.032 | 0.0479* |
Succinate | 0.66 ± 0.037 | 0.0034** |
αKG | 0.80 ± 0.15 | 0.2693 |
Citrate-isocitrate | 0.86 ± 0.07 | 0.0442* |
Aconitase | 1.08 ± 0.026 | 0.1025 |
To confirm the importance of Ca2+ in the regulation of αKGDH activity and its pivotal role in cell survival in cells with defective OXPHOS, we mutated aspartic acid at residue 154 to alanine (D154A) in the Ca2+-binding site of the E1-αKGDH subunit. This mutation decreases the Ca2+ sensitivity of αKGDH activity by six-fold (41). We hypothesized that overexpression of the E1-αKGDH mutant would decrease αKGDH activity and consequently reduce cell viability, mimicking the effects observed after the inhibition of Ca2+ transfer to mitochondria. As a control, we overexpressed E179A E1-αKGDH (fig. S2E), which has a Ca2+ sensitivity similar to the wild-type enzyme. All expressed proteins localized to the mitochondria (fig. S2F). Viability did not differ between overexpressing cells grown in glucose-rich culture media (Fig. 3G). However, when cells were cultured for 24 hours with glutamine as the only carbon source, the cells expressing the D154A mutant had significantly decreased viability, whereas the E179A-expressing and wild-type E1-αKGDH–expressing cells were unaffected (Fig. 3G). This result confirms that mitochondrial Ca2+ is critical for promoting sufficient αKGDH activity to maintain cell viability, and it complements the results achieved by genetic and pharmacological inhibition of mitochondrial Ca2+ uptake.
143BΔcytb cells had a ~ 50% lower NAD+/NADH ratio than 143Bwt cells (fig. S4A), as previously reported (16). Exposure to XeB increased the NAD+/NADH ratio by twofold at 1 hour in both cell lines, which further increased by fivefold in 143BΔcytb cells at 4 hours (Fig. 3H). By comparison, the mitochondrial uncoupler carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP) (1 μM, 1 hour) caused a twofold increase in both cell lines (Fig. 3H). Similar results were observed in RCC4 cells (fig. S2G). To determine whether the observed NAD+/NADH ratio change in response to XeB was localized to mitochondria, we used two-photon microscopy to determine the fluorescence intensity ratio between flavin adenine dinucleotide (FAD) and pyridine nucleotides [NAD(P)H]. This optical redox ratio correlates with the oxidation-reduction state (42, 43) and NAD+/NADH ratio (44). A 4-hour XeB treatment significantly increased the optical redox ratio (Fig. 3, I and J), confirming that NAD+/NADH changed in mitochondria where αKGDH is located. To determine whether there was also a change in cytoplasmic NAD+/NADH, we transfected cells with the NAD+/NADH sensor peredox (45). XeB significantly increased the NAD+/NADH ratio, presented here as the difference between the red-to-green fluorescence signal at the beginning and end of the experiment (fig. S4, B and C). Together, these data confirm that the NAD+/NADH ratio increases in both the mitochondria and cytoplasm when Ca2+ transfer from the ER to mitochondria is inhibited.
The greater XeB-induced increase in the NAD+/NADH ratio observed in the 143BΔcytb compared with 143Bwt cells is consistent with higher basal NADH production through the activity of αKGDH required for reductive carboxylation in the143BΔcytb cells. The XeB-induced increase in the NAD+/NADH ratio is of similar magnitude to that caused by siRNA-mediated knockdown of αKGDH (20). These results further suggest that αKGDH activity depends on mitochondrial matrix Ca2+ supplied by low-level InsP3R-mediated Ca2+ release. To test this, we reasoned that XeB-induced reduction of αKGDH activity and associated cell death could be reversed by supplying sufficient substrates to elevate flux through αKGDH by mass action. We first supplied methyl-pyruvate because it can be converted to acetyl-CoA and citrate, which can be used for reductive metabolism to generate metabolic intermediates needed for growth of the OXPHOS-deficient 143BΔcytb cells. Furthermore, high concentrations of methyl-pyruvate will generate a substrate mass-action effect to accelerate the activity of αKGDH independently of allosteric regulation by Ca2+. Addition of methyl-pyruvate to the growth medium prevented XeB-induced cell death in both cell lines. Dimethyl-α-ketoglutarate rescues other OXPHOS-defective cells that rely on glutamine-derived anapleurosis in a manner dependent on αKGDH activity (17). Similarly, addition of dimethyl-α-ketoglutarate to enhance αKGDH activity by mass action also markedly reduced XeB-induced death in the 143BΔcytb cells (Fig. 3K). To determine whether constitutive ER-to-mitochondrial Ca2+ transfer was generally required by other OXPHOS-deficient cells that rely on reductive carboxylation, we induced reductive carboxylation in 143Bwt and Hela cells by impairing OXPHOS by treatment with antimycin A (16). Antimycin A reduced OCR and elevated ECAR (fig. S4, D to G) similar to the 143BΔcytb and RCC4 cells, and the treated cells became more sensitive to XeB-induced cytotoxicity (fig. S4, H and I). Antimycin A–treated Hela cells also exhibited enhanced cytotoxicity in response to AOA and CPI-613, similar to 143BΔcytb cells (fig. S4, J and K). Together, these data suggest that αKGDH maintains metabolic homeostasis in OXPHOS-defective cancer cells by enabling the use of glutamine for reductive carboxylation and that this activity depends on constitutive ER-to-mitochondria Ca2+ transfer.
Inhibition of InsP3R-mediated Ca2+ release in OXPHOS-defective cells activates SIRT1-mediated prosurvival autophagy
Inhibition of InsP3R activity and Ca2+ uptake by mitochondria induces prosurvival autophagy in several cell lines (23, 24). Inhibition of InsP3R activity with XeB or by simultaneous RNA interference knockdown of types 1 and 3 InsP3R (Fig. 4, A and B) or knockdown of MCU (Fig. 4C) induced autophagy in both 143Bwt and 143BΔcytb cells. Similar results were observed in RCC4 + vhl and RCC4 cells (fig. S2H). Autophagy plays a prosurvival role in both cell lines because exposure to XeB in the presence of the autophagy inhibitors 3-methyladenine (3MA) (Fig. 4D) or chloroquine (CQ; fig. S5A) caused profound cell death (Fig. 4E).
Fig. 4. Inhibition of constitutive InsP3R-mediated Ca2+ release increases SIRT1-dependent deacetylation and autophagy.
(A) Representative Western blots of the autophagy marker LC3 or tubulin as loading control in 143Bwt (wt) and 143BΔcytb (Δc) cell lines treated with 5 μM XeB for 4 hours. (B) As in (A) but with cells transfected with siRNA against types 1 and 3 InsP3R (IP3R1 + 3) or a nontargeting siRNA for 24 hours. (C) As in (A) but with cells transfected with a siRNA against MCU or a nontargeting siRNA for 72 hours. Bar graphs in (A), (B), and (C) represent quantification of LC3-II/tubulin expressed as means ± SEM of three independent experiments. (D) Cell death as assessed by LDH release in cells treated with 5 μM XeB and/or 3-methyladenine (3MA) or vehicle control (C) for 12 hours. (E) As in (D) but with cells transiently transfected with a siRNA against MCU, types 1 and 3 InsP3R (IP3R1 + 3) or a nontargeting siRNA and treated or not with 3MA for 12 hours. Cytotoxicity data in (D) and (E) are means ± SEM of three independent experiments, each with six replicates. (F) Representative Western blots of four independent experiments to determine pan-acetylated lysine or tubulin in cells under basal conditions (C) or pretreated with 10 μM of the SIRT1 inhibitor EX527 (EX) for 12 hours and then exposed to 5 μM XeB for 4 hours. (G) Representative Western blots of LC3 and tubulin in 143Bwt cells under basal conditions (C) exposed or not to 10 μM EX527 (EX) for 12 hours and then treated with 5 μM XeB for 4 hours. (H) As in (G) in 143BΔcytb cells. Bar graphs in (G) and (H) represent quantification of LC3-II/tubulin expressed as means ± SEM of three independent experiments. (I) Representative immunofluorescence images of three independent experiments showing LC3 puncta formation (arrows) in cells under basal conditions (control) or pretreated with 10 μM EX527 for 12 hours and then exposed to 5 μM XeB for 4 hours. Scale bar, 10 μm. (J) Cell death as assessed by LDH release in cells under basal conditions (C) or exposed to 10 μM EX527 (EX) for 12 hours and then treated with 5 μM XeB for 4 hours. Results represent means ± SEM of three independent experiments. For all experiments, *P < 0.05, **P < 0.01, and ***P < 0.001 compared to respective control.
In normal cells and cancer cells with functional mitochondria, inhibition of Ca2+ transfer to mitochondria reduces ATP levels and activates AMPK in addition to activating autophagy (23, 24). Unexpectedly, inhibition and/or knockdown of InsP3R or MCU activated AMPK in both 143Bwt and 143BΔcytb cells (fig. S5, B to D). This activation was associated with an increased AMP/ATP ratio in the 143Bwt cells (fig. S5, E to G), whereas the AMP/ATP ratio was unchanged in the 143BΔcytb cells (fig. S5, E to G), as expected, because ATP levels in these cells are independent of OXPHOS. Thus, mechanisms other than a fall in ATP must be responsible for enhanced AMPK activity observed in the 143BΔcytb cells. Because NAD+ levels are increased by inhibition of Ca2+ transfer to mitochondria (Fig. 3, H to J) and autophagy can be regulated by the NAD+-dependent deacetylase SIRT1 (31), we examined the possible involvement of SIRT1 in XeB-induced autophagy. Global acetylation levels were depressed by XeB, as shown by a reduction in the acetylation of a 53-kDa protein (Fig. 4F), a molecular weight that corresponds to p53, a known SIRT1 substrate (46), which was confirmed by Western blot (fig. S5, H and I). The SIRT1 inhibitor EX527 suppressed the reduction in p53 acetylation caused by XeB (Fig. 4F and fig. S5, H and I). EX527 reduced XeB-induced autophagy in both 143Bwt and 143BΔcytb cell lines, as shown by Western blot (Fig. 4, G and H) and immunofluorescence (Fig. 4I) without altering AMPK phosphorylation (fig. S5, J and K). The levels of SIRT1 were similar in both 143Bwt and 143BΔcytb cell lines (fig. S6A), and SIRT1 was located in both the cytoplasm and nucleus to a similar extent (fig. S6B). To confirm a role of SIRT1 in XeB-induced autophagy, we treated mouse embryonic fibroblasts (MEFs) lacking SIRT1 (SIRT1−/− cells) (fig. S6C) with XeB and monitored the formation of autophagosomes with mCherry–green fluorescent protein (GFP)–LC3. SIRT1−/− cells were unable to mount an autophagic response, as evidenced by the absence of autophagosomal-puncta formation (fig. S6, D and E). In contrast, autophagosome formation was significantly enhanced in SIRT1-expressing wild-type MEFs, confirming that SIRT1 was necessary for XeB-induced autophagy. Inhibition of SIRT1 activity by EX527 increased cell death in both 143Bwt and 143BΔcytb cell lines (Fig. 4J), as well as in RCC4 cells (fig. S2I). These results suggest that elevated NAD+ levels activate SIRT1 to induce prosurvival autophagy in response to inhibition of ER-to-mitochondrial Ca2+ transfer.
Whereas inhibition of constitutive Ca2+ transfer from ER to mitochondria activated prosurvival autophagy in both 143wt and 143BΔcytb cells, 143BΔcytb cells were protected to a lesser extent (Fig. 4, A to E). To understand the basis for the enhanced cytotoxicity observed in the 143BΔcytb cells, we measured autophagic flux in both cell lines. Both lines were treated with CQ, which inhibits fusion of autophagosomes with lysosomes (47). Under these conditions, autophagosomal components, including LC3-II, cannot be degraded and accumulate. CQ caused LC3-II to accumulate to similar extents in both cell lines (Fig. 5A). XeB caused a strong and additive increase in LC3-II accumulation in CQ-treated 143Bwt cells (Fig. 5A), indicating enhanced autophagic flux. In contrast, the CQ- and XeB-induced increases were not additive in 143BΔcytb cells (Fig. 5A), suggesting that autophagic flux was compromised. To confirm this interpretation, we used tandem-tagged mCherry-GFP-LC3. When autophagosomes containing mCherry-GFP-LC3 fuse to lysosomes, GFP fluorescence is quenched, whereas the red fluorescence from mCherry remains. Observation of a mix of yellow and red puncta is indicative of enhanced autophagic flux (48), whereas defective lysosomal fusion results in the accumulation of yellow puncta. Inhibition of InsP3R-mediated Ca2+ release by XeB induced the appearance of a mix of yellow and red puncta in 143Bwt cells (Fig. 5, B and C). In contrast, 143BΔcytb cells accumulated only yellow puncta (Fig. 5, B and C). This result suggests that inhibition of constitutive Ca2+ release enhances autophagic flux in cells with normal OXPHOS, as previously reported, whereas there is a defect in autophagosome-lysosome fusion in 143BΔcytb cells. To determine whether defective autophagosome-lysosome fusion was specific to 143BΔcytb cells or was a consequence of their metabolic reliance on reductive carboxylation, reductive carboxylation was induced in Hela cells by treatment with antimycin A. As in the 143BΔcytb cells, XeB-induced autophagy in the antimycin A–treated Hela cells caused an accumulation of yellow autophagosomal puncta (fig. S6, F and G). We speculated that because autophagy provides necessary metabolites to maintain cellular homeostasis (49), impaired autophagic flux in cells with defective OXPHOS in response to decreased Ca2+ transfer to mitochondria likely increases metabolic stress that results in the observed enhanced cytotoxicity in these cells. Cells with defective OXPHOS are deficient in aspartate synthesis necessary for nucleoside synthesis (18, 19). Addition of nucleosides as a metabolic substrate to the media prevented XeB-induced cell death in both cell lines (Fig. 5D).
Fig. 5. Autophagy flux is impaired in OXPHOS-defective cells.
(A) Representative Western blots of autophagy marker LC3 and tubulin as loading control in 143Bwt (wt) and 143BΔcytb (Δc) cell lines treated with 5 μM XeB for 4 hours in the presence of the autophagy inhibitor chloroquine (CQ). Bar graphs represent quantification of LC3-II/tubulin expressed as means ± SEM of three independent experiments. ***P < 0.001 compared to respective control. (B) Representative images of 143Bwt (wt) and 143BΔcytb (Δc) cell lines transfected with mCherry-GFP-LC3 for 24 hours, followed by exposure to 5 μM XeB for 4 hours. Arrows indicate autophagosome (yellow) and autolysosome (red). Scale bar, 10 μm. (C) Number of autolysosomes (red) and autophagosome (yellow) per cell. Results represent means ± SEM of three independent experiments. At least 100 cells per experiment were counted. ***P < 0.001 compared to respective control. (D) Cell death by LDH release in cells under basal conditions (C) or exposed to 5 μM XeB for 4 hours in the presence of nucleosides (Nx). Means ± SEM of three independent experiments. ***P < 0.001 compared to respective control.
DISCUSSION
Whereas the role of mitochondrial Ca2+ in regulating OXPHOS is well established, its role in regulating other mitochondrial metabolic pathways is less clearly understood. Here, we explored the role of ER-to-mitochondrial Ca2+ transfer on metabolism and viability of cells with defective capacity for OXPHOS. We found that despite impaired OXPHOS, mitochondrial matrix Ca2+ remained a critical regulator of mitochondrial metabolism and viability because of the Ca2+ sensitivity of αKGDH activity. In addition, we found that the bioenergetic crisis induced by interruption of mitochondrial Ca2+ uptake was sensed by SIRT1, which played a prosurvival role by triggering NAD+-dependent autophagy that was not fully functional in cells with defective OXPHOS. Our results suggest that mitochondrial matrix Ca2+ plays a fundamental role in cellular bioenergetics by regulating both OXPHOS-dependent and OXPHOS-independent metabolic pathways. The discovery of a role for Ca2+ in regulating OXPHOS-independent pathways may have consequences for cancer cell metabolism and possible therapeutic implications.
Matrix Ca2+ concentration regulates mitochondrial bioenergetics, in part, by modulating the activities of the respiratory chain, PDH and two TCA cycle dehydrogenases, NAD+-IDH, and αKGDH. In cells with functional mitochondria, constitutive ER-to-mitochondrial transfer of Ca2+ supports the activity of PDH as a key enzyme that controls carbon flux into the TCA cycle. In the absence of this transfer, Ca2+-activated PDH phosphatase activity decreases, resulting in PDH hyperphosphorylation and diminished activity, limiting flux of pyruvate into the TCA cycle. αKGDH oxidizes αKG to succinyl-CoA, CO2, and NADH, providing reducing equivalents for OXPHOS (50). Diminished TCA cycle production of NADH leads to lower rates of OXPHOS, and ATP levels decline (23). Tumorigenic breast, prostate, and melanoma cancer cells behave similarly (24), supporting the concept that many cancer cells also rely on mitochondrial metabolism for ATP generation (1, 2, 51, 52). Whether matrix Ca2+ plays an important role in mitochondrial bioenergetics beyond its role in regulating OXPHOS has been unknown. To explore the role of matrix Ca2+ in regulating mitochondrial pathways in addition to OXPHOS, we examined 143BΔcytb cells as a model for cells that have defective OXPHOS. αKGDH plays a central role in these cells as a gatekeeper for utilization of glutamine-derived carbon for oxidation and reductive carboxylation, with its importance in cancer cell metabolism well recognized (6, 7, 20, 40, 51). Many cancers display oncogene-dependent addiction to glutamine that has been exploited therapeutically (53). Through glutaminase activity, glutamine provides glutamate that is converted by glutamate dehydrogenases or aminotransferases to αKG (2). When oxidation of metabolites in the TCA cycle is unfavorable, such as in the presence of defective OXPHOS activity or hypoxia, many cancer cells use glutamine-derived αKG through reductive carboxylation to generate citrate for lipid synthesis and four-carbon intermediates needed to produce other TCA cycle metabolites (40). The oxidation of αKG through αKGDH in these circumstances remains essential because it is necessary to generate the reducing equivalents necessary to maintain reductive carboxylation (20). Of the Ca2+-sensitive enzymes in the TCA cycle, only αKGDH contributes to generation of metabolites in cells performing reductive carboxylation, including 143BΔcytb cells (16). αKGDH inhibition impairs the viability of several cancer cell lines (54, 55) and breast cancer–associated lung metastasis (56).
We found that inhibition of InsP3R-mediated Ca2+ release or mitochondrial Ca2+ uptake decreased αKGDH activity in both OXPHOS-competent 143Bwt cells and OXPHOS-defective 143BΔcytb cells, as evidenced by αKG accumulation and an increased NAD+/NADH ratio. Similar accumulation of αKG suggests that αKGDH activity is comparable in both cell lines, in accordance with the similar glutamine consumption rates in these cells (16). Accumulation of αKG in response to reduced mitochondrial Ca2+ uptake in 143Bwt cells is consistent with the Ca2+ dependence of αKGDH activity. Reductive carboxylation of αKG is mediated by NADP-IDH2, which is not Ca2+ sensitive (57). Thus, αKG accumulation observed in 143BΔcytb cells is likely caused by restricted availability of reducing equivalents, which are normally generated by Ca2+-dependent αKGDH oxidation of αKG and are necessary to fuel IDH2-mediated reductive carboxylation of αKG. Inhibition of InsP3R activity induced a larger increase in the NAD+/NADH ratio in 143BΔcytb cells compared with 143Bwt cells, consistent with a higher rate of NADH consumption in 143BΔcytb cells to sustain their higher rates of reductive carboxylation (20).
As shown previously, inhibition of InsP3R Ca2+ signaling and mitochondrial Ca2+ uptake induces a bioenergetic crisis that results in activation of prosurvival autophagy. Similar responses were observed for the OXPHOS-competent and OXPHOS-incompetent cells used here. Previously, we demonstrated that activation of AMPK, due to an elevated adenosine diphosphate/ATP ratio, is upstream of autophagy activation (23). Here, we found that inhibition of constitutive ER-to-mitochondria Ca2+ transfer was associated with an elevated NAD+/NADH ratio that was sensed by SIRT1. Pharmacological and genetic inhibition of SIRT1 repressed autophagy, indicating that SIRT1 was also upstream of autophagy activation. Prolonged inhibition of mitochondrial Ca2+ uptake reduced the viability of both 143Bwt and 143BΔcytb cells. Cell death was prevented by incubation with cell-permeable methyl-pyruvate, dimethyl-αKG, and nucleosides, suggesting that it was caused by deficits in mitochondrial bioenergetics. The relatively minor cytotoxicity observed in 143Bwt cells was likely due to an effective AMPK-dependent (23) and SIRT1-dependent prosurvival autophagic response to inhibition of OXPHOS caused by reduced activities of PDH and TCA cycle dehydrogenases (24). The enhanced death rate of 143BΔcytb cells may be accounted for by two, not mutually exclusive mechanisms. First, the combination of OXPHOS inhibition (reduced ATP and ψm) and lack of essential TCA cycle–derived metabolic intermediates derived from reductive carboxylation, due to the central role of Ca2+-dependent αKGDH activity in this process, may critically compromise cell viability. We considered that the higher sensitivity of 143BΔcytb cells could be the result of excessive ROS production, but we observed no differences between the two cell lines, in agreement with other reports (6, 38), and the antioxidant NAC did not confer protection. In contrast, the two cell lines differed in their autophagic responses to their distinct metabolic challenges. Whereas increased autophagic flux in 143Bwt cells culminated in the delivery of autophagosomal cargo to lysosomes, autophagic flux in the 143BΔcytb cells was aborted, leaving autophagosome accumulation in the cytoplasm as a consequence of a failure to fuse with lysosomes. Because autophagy as a survival mechanism requires lysosomal-mediated hydrolysis reactions to recycle metabolic intermediates for use elsewhere in the cell, diminished autophagosome-lysosome fusion is expected to limit the availability of lysosome-derived metabolites necessary for cell death protection. The lack of both lysosome-derived metabolic intermediates and TCA-derived metabolites together may create conditions that are highly cytotoxic in OXPHOS-defective cells. The mechanisms that limit autophagosome-lysosome fusion specifically in 143BΔcytb cells remain to be determined. Alterations in mitochondrial function caused by chemical inhibition of the electron transport chain impairs lysosomal activity and induces accumulation of large vacuoles (58), but this seems unlikely to account for our observations because those effects are mediated by ROS, which were produced at similar levels in 143Bwt and 143BΔcytb cells. CD4+ T lymphocytes with defective mitochondrial function caused by genetic deletion of the mitochondrial transcription factor Tfam show dysfunctional lysosomes and impaired endolysosomal trafficking and autophagy. However, in contrast to our results, those phenotypes are associated with a reduced NAD+/NADH ratio, and restoration of NAD+ levels improves lysosomal function (59). However, these results do suggest that cell redox state may fine-tune lysosomal function and play a role in 143BΔcytb cells.
In summary, the results in the present study show that inhibition of constitutive mitochondrial Ca2+ uptake elevates NAD+ by reducing αKGDH activity, which promotes SIRT1 activation that, in turn, activates autophagy as a prosurvival mechanism. In cells with defective OXPHOS, impaired autophagosome-lysosome fusion exacerbates metabolic stress, rendering these cells more sensitive to cell death. Our results highlight a central role of αKGDH and OXPHOS in cancer cells and suggest that the mechanisms involved in ER-to-mitochondria Ca2+ transfer offer new therapeutic targets for cancer.
MATERIALS AND METHODS
Reagents
The LC3B (#2775), acetylated-lysine (#9441), acetyl-p53 (#2525), phospho-AMPK (#2535), AMPK (#5831), SIRT1 (#2493), and p53 (#2527) antibodies were from Cell Signaling Technology. The β-tubulin (#556321) and InsP3R-3 (#610313) antibodies were from BD Biosciences. The MCU (#HPA016480) and c-Myc (#M4439) antibodies were from Sigma-Aldrich. The InsP3R-1 (ABS55) antibody was from Millipore. Anti-OGDH (2-oxoglutarate dehydrogenase) rabbit polyclonal (#GTX33374) was from GeneTex. Nuclei were stained with Hoechst. Peroxidase-conjugated secondary antibodies were purchased from Amersham (#NA934 and #NA931). 3MA, methyl-pyruvate, dimethyl-α-ketoglutarate, antimycin A, FCCP, oligomycin, rotenone, 2DG, and Hank’s balanced salt solution (HBSS) were from Sigma-Aldrich. CPI-613 and EX527 were from Tocris Biosciences. Fluo-4 AM, NAO, and Hoechst were from Thermo Fisher Scientific. EGCG and CB839 were from Selleckchem. XeB was extracted and purified from the marine sponge Xestospongia exigua as described (32).
Cell culture and transfection
Osteosarcoma 143Bwt and 143BΔcytb cybrid cells, provided by N. S. Chandel (Northwestern University), were cultured in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal bovine serum (FBS), uridine (100 μg/ml), 1 mM pyruvate, penicillin (100 U ml−1), streptomycin (100 μg ml−1), and fungizone (0.25 μg ml−1; Gibco) at 37°C (95%/5% air/CO2). Hela cells were grown in similar conditions in the absence of uridine and pyruvate. Wild-type and SIRT1−/− cells, provided by Leonard Guarente (Massachusetts Institute of Technology), were cultured in similar conditions as 143B cells. Pharmacological inhibition of OXPHOS in 143Bwt and Hela cells was achieved by adding 1 μM antimycin A (complex III inhibitor) to the culture media for 48 hours, as described (16). Transfections of siRNA were performed with a Nucleofector electroporator (Amaxa Biosystems). siRNAs against MCU and types 1 and 3 InsP3Rs were previously described (23) and purchased from GE Healthcare Dharmacon. To simultaneously silence type 1 and type 3 InsP3R, a specific siRNA was designed (GE Healthcare Dharmacon), which consisted of 5′-GUGAGAAGCAGAAGAAGGAUU-3′ and 3′-UUCACUCUUCGUCUUCUUCCU-5′. pAAV-CAG-Peredox was a gift from G. Yellen (Addgene plasmid # 73807; http://n2t.net/addgene:73807; RRID:Addgene_73807).
Ca2+ signals
Imaging of spontaneous changes in cytoplasmic Ca2+ concentration ([Ca2+]c) was accomplished by total internal reflection fluorescence (TIRF) microscopy using a Nikon eclipse Ti inverted TIRF microscope equipped with an apo TIRF 60×, 1.49 numerical aperture (NA) lens with a stage-top Tokai Hit incubator. Cells were loaded with freshly prepared Fluo-4 (5 μM) and imaged at 37°C and 5% CO2 ~100 nm into the cell. The gain on the Andor iXon EMCCD camera was set to maximum, and images were collected every 50 ms for 2 min. For simultaneous measurements of [Ca2+]c and mitochondrial matrix [Ca2+] ([Ca2+]m), 143B cells were grown on 25-mm glass coverslips for 48 hours and then loaded with 2 μM rhod-2/AM (50 min) and 5 μM Fluo-4/AM (30 min) in extracellular medium as described (25). Coverslips were mounted in an open chamber and imaged. After 25 s of baseline recording, an InsP3 mobilizing agonist (histamine, 100 μM) was added and confocal images were recorded every 3 s (510 Meta; Carl Zeiss) at 488- and 561-nm excitation using a ×63 oil objective. Images were analyzed and quantified using ImageJ (National Institutes of Health).
Site-directed mutagenesis
Mutagenesis of recombinant human OGDH cDNA in the Myc-DDK–tagged vector was performed using the QuikChange site-directed mutagenesis kit (Stratagene, La Jolla, CA) as per the manufacturer’s instructions and as previously described (60). For the OGDHD154A-myc mutant, the primers were 5′-CCC TGG GGA TTT TGG CTG CTG ATC TGG ACT C-3′ and 5′-GAC TCC AGA TCA GCA GCC AAA ATC CCC AGG G-3′. For OGDHE179A-myc, the primers were 5′-TTC TAT GGC CTG GAT GCG TCT GAC CTC GAC AAG-3′ and 5′-CTT GTC GAG GTC AGA CGC ATC CAG GCC ATA GAA-3′. All mutants were checked by DNA sequencing, and their expression was analyzed by immunoblot and immunofluorescence assays.
Western blotting and treatments
Drugs were added in fresh media as indicated. Treatments were terminated by rapid removal of medium with cells on ice, followed by cell lysis with CytoBuster protein extraction reagent (Novagen) supplemented with protease and phosphatase inhibitors (cOmplete PhosSTOP, Roche). Protein extracts were separated in 4, 10, or 15% SDS-polyacrylamide gels and transferred to polyvinylidene difluoride membranes (Millipore). Blocking was performed at room temperature for 1 hour in 5% fat-free milk, and membranes were incubated overnight at 4°C with primary antibody and then for 1 hour at room temperature with a secondary antibody conjugated to horseradish peroxidase. Chemiluminescence detection used ECL Plus reagent (Pierce) and a series of timed exposure images were acquired with a FluorChem Q system (ProteinSimple) to ensure that densitometric analyses were performed at exposures within the linear range. To ensure equal protein loading across gels, membranes were submerged in stripping buffer (Restore Western Blot Stripping Buffer; Pierce), incubated at 37°C for 20 min, and reprobed with a loading control antibody. ImageJ was used for densitometric analysis.
Cellular oxygen consumption and ECARs
OCR and ECAR as measurements of OXPHOS and glycolysis, respectively, were measured at 37°C using an XFe96 extracellular analyzer (Agilent, USA). A total of 1.5 × 104 cells per well were seeded onto poly-lysine (Sigma-Aldrich, USA) pretreated plates and allowed to attach for 24 hours. Where indicated, 143Bwt and 143BΔcytb cells were exposed either to 5 μM XeB or vehicle for 4 hours and then loaded into the analyzer in fresh unbuffered Seahorse media, and basal OCR and ECAR were determined. Sequential injections of 1 μM oligomycin, 50 nM FCCP, and 1 μM rotenone/antimycin A were used to reveal different parameters of cellular respiration.
Immunofluorescence
Cells were fixed in 100% methanol at −20°C for 12 min, blocked in 1% bovine serum albumin, and incubated overnight with LC3B antibody (Cell Signaling technology, #2775). Alexa Fluor 488 secondary antibody (Molecular Probes, #A-11008) was incubated for 1 hour at room temperature, and cells were mounted in ProLong Gold anti-fade with DAPI (4′,6-diamidino-2-phenylindole; Molecular Probes, #P36931). Visual examination was performed on a Zeiss Axiovert 510 LSM Pascal confocal microscope using a high–numerical aperture water immersion 63× objective.
Optical redox ratio calculation
Microscopy images were collected using a Leica TCS SP8 confocal microscope system equipped with an Insight DS+ Ti:Sapphire laser (Spectra-Physics) for multiphoton near-infrared (NIR) excitation. Two-photon excitation fluorescence images were obtained at excitation wavelengths of 755 and 860 nm, with average powers of 18 and 29 mW at the sample, respectively. Two emission channels were recorded simultaneously on a pair of nondescanned Leica HyD hybrid detectors, using a fluorescein isothiocyanate/DAPI fluorescence filter cube equipped with a 680-nm short-pass entrance filter and emission band-pass filters at 460 ± 25 nm (NADH channel) and 525 ± 25 nm (FAD channel). A Leica DMi8 inverted microscope was used with a Leica HC PL IRAPO 40×/1.10 water immersion objective optimized for confocal NIR microscopy. Images were collected with a 1.5× zoom magnification factor and sampled using 8-bit dynamic range and 1024 × 1024 pixels (corresponding to a 194 μm by 194 μm field of view, or 0.189 μm/pixel). Bidirectional raster scans were conducted at a line rate of 600 Hz and averaging each line eight times per image. For each of the four types of samples studied (ΔcytB and wt cells, both XeB treated and untreated), images were collected from three wells in a multiwell glass-bottomed dish, and sampling between three to five different regions of interest within each well. Fluorescence intensity images of NAD(P)H and FAD were normalized on the basis of the illumination power and averaging. The background and weak nucleus fluorescence were removed using a binary mask with Otsu’s threshold method to exclude noise from signal not pertaining to metabolism. After co-registering the fluorescence intensity images using cross-correlation, the pixel-wise optical ratio [FAD/(FAD + NAD(P)H] was calculated and the mean value per field was compared (61).
Peredox
Peredox-mCherry was transfected into cells using FuGene according to the manufacturer’s protocol and imaged in live cells 12 hours later. Peredox is composed of a permuted GFP T-Sapphire (green) interposed between two Rex subunits, with a C-terminal mCherry to normalize for green fluorescence (45). Ratiometric imaging (green/red) was performed on a Nikon A1R confocal microscope equipped with Perfect Focus using a 63× plan apo lens/1.4 NA. Peredox increases GFP fluorescence (excitation, 405; emission 520) upon binding to NADH. The green-to-red fluorescence ratio increase, which correlates with an increase in the NADH:NAD+ ratio, was analyzed using SML software and ImageJ.
Mitochondrial membrane potential
Cells were loaded with tetramethylrhodamine methyl ester (TMRM; Life Technologies; 5 nM) for 30 min at 37°C and 5% CO2. Hoechst 33342 was used as a nuclear counterstain. Images were acquired using a Nikon C2+ spectral confocal microscope using the same settings for all experiments. TMRM fluorescence was quantified with ImageJ software.
Determination of ROS levels
The generation of intracellular oxidative stress was determined using the dihydroethidium (DHE) probe as previously described (62). 143Bwt and 143BΔcytb cells were grown in complete media, seeded in 12-well plates, and allowed to attach overnight. The next day, the cells were exposed for 4 hours to ethanol (control) or 5 μM XeB, and then, the culture medium was replaced by a solution containing 5 μM DHE in HBSS and incubated for 20 min in the dark. After this, the cells were washed, trypsinized, and resuspended in 500 μl of HBSS and measured by FACS Calibur flow cytometer.
Determination of NAD+/NADH ratio and αKG levels
All measurements were performed using commercial kits (BioVision, K337–100; K677–100) according to the manufacturer’s instructions. Briefly, cells were incubated either with 5 μM XeB or vehicle for 1 or 4 hours, washed twice with phosphate-buffered saline (PBS), scraped off the dishes, and pelleted. For NAD+/NADH ratio, NAD+ and NADH were extracted and the samples were subjected to two freeze-thaw cycles and centrifuged. Aliquots of each sample were heated at 60°C for 30 min to decompose NAD+. Samples were then loaded into 96-well plates for absorbance measurements at 450 nm. To determine αKG levels, samples were deproteinized with 7% perchloric acid, neutralized with a buffer solution (3 M KOH, 0.4 M tris, and 3 M KCl), and converted to pyruvate. Background pyruvate was subtracted. The reaction was incubated at 37°C for 30 min, and optical density was measured at 570 nm.
Metabolomics
Subconfluent culture dishes were incubated for 4 hours in DMEM containing 15 mM glucose, 2 mM glutamine, and 10% FBS in the presence or absence of 5 μM XeB. Cells were then washed twice with ice-cold saline and overlaid with 500 μl of cold methanol/water (50/50, v/v), transferred to an Eppendorf tube and subjected to three freeze-thaw cycles. After rigorous vortexing, the debris was pelleted by centrifugation at 16,000g and 4°C for 15 min. Pellets were used for protein quantitation (BCA Protein Assay; Thermo Fisher Scientific). The supernatant was transferred to a new tube and evaporated to dryness using a SpeedVac concentrator (Thermo Savant, Holbrook, NY). Metabolites were reconstituted in 100 μl of 0.03% formic acid in analytical-grade water, vortex-mixed, and centrifuged to remove debris. Thereafter, the supernatant was transferred to a high-performance liquid chromatography (HPLC) vial for the metabolomics study. Targeted metabolite profiling was performed using a liquid chromatography–mass spectrometry/mass spectrometry (LC-MS/MS) approach as described before (20). Briefly, metabolites were reconstituted in 100 μl of 0.03% formic acid in LC-MS–grade water, vortex-mixed, and centrifuged to remove debris. LC-MS/MS and data acquisition were performed using an AB QTRAP 5500 LC/triple quadrupole mass spectrometer (Applied Biosystems SCIEX) with an injection volume of 20 μl. Chromatogram review and peak area integration were performed using MultiQuant software version 2.1 (Applied Biosystems SCIEX). The peak area for each detected metabolite was normalized against the total ion count of each sample and the mean for each metabolite within the sample batch of each run to correct for variations introduced by sample handling through instrument analysis. The normalized areas were used as variables for the multivariate analyses and modeling using SIMCA-P (version 13.0.1; Umetrics). The processed datasets were mean-centered, unit-variance–scaled, and then applied to principal components analysis to evaluate the clustering and to detect outliers. Univariate statistical differences of the metabolites between two groups were analyzed using Student’s t test. The pathway enrichment analysis of differential metabolites between two groups was performed using Metaboanalyst 3.0 (http://metaboanalyst.ca).
Determination of intracellular ATP levels
ATP levels were determined with CellTiter-Glo Luminescent Cell Viability Assay kit (Promega, USA). 143Bwt and 143BΔcytb cells (1 × 105 cells/ml) were seeded into 96-well plates and incubated for 4 hours in culture medium in the absence (vehicle, ethanol) or presence of 5 μM XeB, 1 μM antimycin A, and 1 μM oligomycin. After exposure, cells were washed twice with PBS to remove the medium and resuspended in 20 μl of PBS. Bioluminescence was measured on a BioTek luminometer.
AMP/ATP ratio
Nucleotides were extracted using perchloric acid, neutralized, and frozen for subsequent HPLC analysis. AMP and ATP were quantified using ion pair reverse-phase HPLC, with a C18 RP column, under isocratic elution conditions in 200 mM phosphate, 5 mM tetra-butylammonium phosphate, and 3% acetonitrile.
Cell death assays
Lactate dehydrogenase (LDH) released from dying cells was determined by a colorimetric assay (Roche). Propidium iodide incorporation (5 mg/ml) was determined by flow cytometry (BD FACSAria III).
Statistics
All statistical analyses were performed using GraphPad Prism 4.03 (GraphPad Software, San Diego, CA, USA). The data are expressed as means ± SEM of three independent experiments, each one performed in technical triplicate. Statistical analysis was performed using unpaired t tests, one-way analysis of variance (ANOVA) with Bonferroni’s post-test for pairwise comparisons, or two-way ANOVA. The data were considered statistically significant at the 95% level (P < 0.05).
Supplementary Material
Fig. S1. Mitochondrial parameters and simultaneous knockdown of types 1 and 3 InsP3R induces cell death in OXPHOS-defective 143BΔcytb cells.
Fig. S2. Pseudohypoxic RCC4 cells rely on constitutive InsP3R-mediated Ca2+ signals.
Fig. S3. Inhibition of constitutive InsP3R-mediated Ca2+ transfer to the mitochondria substantially changes the metabolomic profile of 143BΔcytb cell lines.
Fig. S4. OXPHOS-defective cells rely on constitutive InsP3R-mediated Ca2+ signals.
Fig. S5. Inhibition of constitutive InsP3R-mediated Ca2+ release reduces αKGDH activity and activates SIRT1 and AMPK.
Fig. S6. SIRT1 is necessary for the autophagy induced by inhibition of Ca2+ transfer to mitochondria.
Acknowledgments:
We thank C. Moraes (University of Miami) and I. F. M. de Coo (Maastricht University) for generating cytochrome b mutant cells, N. S. Chandel (Northwestern University) for providing them, and F. Gonzalez for assistance in the statistical analyses of the data.
Funding:
This work was supported by FONDECYT postdoctoral fellowship nos. 3140458 (F.J.) and 3170813 (F.U.), NIH P30NS047243 (A.L.), FONDECYT nos. 1160332 (C.C.) and 11170291 (F.J.), ANID/FONDAP no. 15150012 (C.C.), NIH R37GM56328 (J.K.F.), NIH S10OD021624 (I.G.), and a grant from the Emerson Collective Cancer Research Fund (J.K.F.).
Footnotes
SUPPLEMENTARY MATERIALS
Competing interests: The authors declare that they have no competing interests.
Data and materials availability:
All data needed to evaluate the conclusions in the paper are present in the paper or the Supplementary Materials.
REFERENCES AND NOTES
- 1.Koppenol WH, Bounds PL, Dang CV, Otto Warburg’s contributions to current concepts of cancer metabolism. Nat. Rev. Cancer 11, 325–337 (2011). [DOI] [PubMed] [Google Scholar]
- 2.Pavlova NN, Thompson CB, The emerging hallmarks of cancer metabolism. Cell Metab. 23, 27–47 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Wellen KE, Thompson CB, A two-way street: Reciprocal regulation of metabolism and signalling. Nat. Rev. Mol. Cell Biol 13, 270–276 (2012). [DOI] [PubMed] [Google Scholar]
- 4.Carracedo A, Cantley LC, Pandolfi PP, Cancer metabolism: Fatty acid oxidation in the limelight. Nat. Rev. Cancer 13, 227–232 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Sanderson SM, Locasale JW, Revisiting the Warburg effect: Some tumors hold their breath. Cell Metab. 28, 669–670 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Weinberg F, Hamanaka R, Wheaton WW, Weinberg S, Joseph J, Lopez M, Kalyanaraman B, Mutlu GM, Budinger GRS, Chandel NS, Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc. Natl. Acad. Sci. U.S.A 107, 8788–8793 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Ahn CS, Metallo CM, Mitochondria as biosynthetic factories for cancer proliferation. Cancer Metab. 3, 1 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Jose C, Bellance N, Rossignol R, Choosing between glycolysis and oxidative phosphorylation: A tumor’s dilemma? Biochim. Biophys. Acta 1807, 552–561 (2011). [DOI] [PubMed] [Google Scholar]
- 9.Moreno-Sanchez R, Marín-Hernández A, Saavedra E, Pardo JP, Ralph SJ, Rodríguez-Enríquez S, Who controls the ATP supply in cancer cells? Biochemistry lessons to understand cancer energy metabolism. Int. J. Biochem. Cell Biol 50, 10–23 (2014). [DOI] [PubMed] [Google Scholar]
- 10.Liberti MV, Locasale JW, The Warburg effect: How does it benefit cancer cells? Trends Biochem. Sci 41, 211–218 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Vander Heiden MG, DeBerardinis RJ, Understanding the intersections between metabolism and cancer biology. Cell 168, 657–669 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Ippolito L, Morandi A, Taddei ML, Parri M, Comito G, Iscaro A, Raspollini MR, Magherini F, Rapizzi E, Masquelier J, Muccioli GG, Sonveaux P, Chiarugi P, Giannoni E, Cancer-associated fibroblasts promote prostate cancer malignancy via metabolic rewiring and mitochondrial transfer. Oncogene 38, 5339–5355 (2019). [DOI] [PubMed] [Google Scholar]
- 13.Naguib A, Mathew G, Reczek CR, Watrud K, Ambrico A, Herzka T, Salas IC, Lee MF, El-Amine N, Zheng W, Di Francesco ME, Marszalek JR, Pappin DJ, Chandel NS, Trotman LC, Mitochondrial complex I inhibitors expose a vulnerability for selective killing of Pten-null cells. Cell Rep. 23, 58–67 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Molina JR, Sun Y, Protopopova M, Gera S, Bandi M, Bristow C, McAfoos T, Morlacchi P, Ackroyd J, Agip A-NA, Al-Atrash G, Asara J, Bardenhagen J, Carrillo CC, Carroll C, Chang E, Ciurea S, Cross JB, Czako B, Deem A, Daver N, de Groot JF, Dong J-W, Feng N, Gao G, Gay J, Do MG, Greer J, Giuliani V, Han J, Han L, Henry VK, Hirst J, Huang S, Jiang Y, Kang Z, Khor T, Konoplev S, Lin Y-H, Liu G, Lodi A, Lofton T, Ma H, Mahendra M, Matre P, Mullinax R, Peoples M, Petrocchi A, Rodriguez-Canale J, Serreli R, Shi T, Smith M, Tabe Y, Theroff J, Tiziani S, Xu Q, Zhang Q, Muller F, DePinho RA, Toniatti C, Draetta GF, Heffernan TP, Konopleva M, Jones P, Di Francesco ME, Marszalek JR, An inhibitor of oxidative phosphorylation exploits cancer vulnerability. Nat. Med 24, 1036–1046 (2018). [DOI] [PubMed] [Google Scholar]
- 15.Wheaton WW, Weinberg SE, Hamanaka RB, Soberanes S, Sullivan LB, Anso E, Glasauer A, Dufour E, Mutlu GM, Budigner GS, Chandel NS, Metformin inhibits mitochondrial complex I of cancer cells to reduce tumorigenesis. eLife 3, e02242 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Mullen AR, Wheaton WW, Jin ES, Chen P-H, Sullivan LB, Cheng T, Yang Y, Linehan WM, Chandel NS, DeBerardinis RJ, Reductive carboxylation supports growth in tumour cells with defective mitochondria. Nature 481, 385–388 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Chen Q, Kirk K, Shurubor YI, Zhao D, Arreguin AJ, Shahi I, Valsecchi F, Primiano G, Calder EL, Carelli V, Denton TT, Beal MF, Gross SS, Manfredi G, D’Aurelio M, Rewiring of glutamine metabolism is a bioenergetic adaptation of human cells with mitochondrial DNA mutations. Cell Metab. 27, 1007–1025.e5 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Birsoy K, Wang T, Chen WW, Freinkman E, Abu-Remaileh M, Sabatini DM, An essential role of the mitochondrial electron transport chain in cell proliferation is to enable aspartate synthesis. Cell 162, 540–551 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Sullivan LB, Gui DY, Hosios AM, Bush LN, Freinkman E, Vander Heiden MG, Supporting aspartate biosynthesis is an essential function of respiration in proliferating cells. Cell 162, 552–563 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Mullen AR, Hu Z, Shi X, Jiang L, Boroughs LK, Kovacs Z, Boriack R, Rakheja D, Sullivan LB, Linehan WM, Chandel NS, DeBerardinis RJ, Oxidation of alpha-ketoglutarate is required for reductive carboxylation in cancer cells with mitochondrial defects. Cell Rep. 7, 1679–1690 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Strumilo S, Short-term regulation of the alpha-ketoglutarate dehydrogenase complex by energy-linked and some other effectors. Biochemistry 70, 726–729 (2005). [DOI] [PubMed] [Google Scholar]
- 22.McCormack JG, Denton RM, The effects of calcium ions and adenine nucleotides on the activity of pig heart 2-oxoglutarate dehydrogenase complex. Biochem. J 180, 533–544 (1979). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Cárdenas C, Miller RA, Smith I, Bui T, Molgó J, Müller M, Vais H, Cheung K-H, Yang J, Parker I, Thompson CB, Birnbaum MJ, Hallows KR, Foskett JK, Essential regulation of cell bioenergetics by constitutive InsP3 receptor Ca2+ transfer to mitochondria. Cell 142, 270–283 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Cárdenas C, Müller M, McNeal A, Lovy A, Jaňa F, Bustos G, Urra F, Smith N, Molgó J, Diehl JA, Ridky TW, Foskett JK, Selective vulnerability of cancer cells by inhibition of Ca2+ transfer from endoplasmic reticulum to mitochondria. Cell Rep. 14, 2313–2324 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Mallilankaraman K, Cárdenas C, Doonan PJ, Chandramoorthy HC, Irrinki KM, Golenár T, Csordás G, Madireddi P, Yang J, Müller M, Miller R, Kolesar JE, Molgó J, Kaufman B, Hajnóczky G, Foskett JK, Madesh M, MCUR1 is an essential component of mitochondrial Ca2+ uptake that regulates cellular metabolism. Nat. Cell Biol 14, 1336–1343 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Mallilankaraman K, Doonan P, Cárdenas C, Chandramoorthy HC, Müller M, Miller R, Hoffman NE, Gandhirajan RK, Molgó J, Birnbaum MJ, Rothberg BS, Mak D-OD, Foskett JK, Madesh M, MICU1 is an essential gatekeeper for MCU-mediated mitochondrial Ca2+ uptake that regulates cell survival. Cell 151, 630–644 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Foskett JK, Philipson B, The mitochondrial Ca2+ uniporter complex. J. Mol. Cell. Cardiol 78, 3–8 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Marin-Valencia I, Yang C, Mashimo T, Cho S, Baek H, Yang X-L, Rajagopalan KN, Maddie M, Vemireddy V, Zhao Z, Cai L, Good L, Tu BP, Hatanpaa KJ, Mickey BE, Matés JM, Pascual JM, Maher EA, Malloy CR, DeBerardinis RJ, Bachoo RM, Analysis of tumor metabolism reveals mitochondrial glucose oxidation in genetically diverse human glioblastomas in the mouse brain in vivo. Cell Metab. 15, 827–837 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Kim Y-R, Baek J-I, Kim SH, Kim M-A, Lee B, Ryu N, Kim K-H, Choi D-G, Kim H-M, Murphy MP, Macpherson G, Choo Y-S, Bok J, Lee K-Y, Park J-W, Kim U-K, Therapeutic potential of the mitochondria-targeted antioxidant MitoQ in mitochondrial-ROS induced sensorineural hearing loss caused by Idh2 deficiency. Redox Biol. 20, 544–555 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Rodríguez-Enríquez S, Pacheco-Velázquez SC, Marín-Hernández Á, Gallardo-Pérez JC, Robledo-Cadena DX, Hernández-Reséndiz I, García-García JD, Belmont-Díaz J, López-Marure R, Hernández-Esquivel L, Sánchez-Thomas R, Moreno-Sánchez R, Resveratrol inhibits cancer cell proliferation by impairing oxidative phosphorylation and inducing oxidative stress. Toxicol. Appl. Pharmacol 370, 65–77 (2019). [DOI] [PubMed] [Google Scholar]
- 31.Moreno-Sánchez R, Rodríguez-Enríquez S, Marín-Hernández A, Saavedra E, Energy metabolism in tumor cells. FEBS J. 274, 1393–1418 (2007). [DOI] [PubMed] [Google Scholar]
- 32.Cárdenas C, Liberona JL, Molgó J, Colasante C, Mignery GA, Jaimovich E, Nuclear inositol 1,4,5-trisphosphate receptors regulate local Ca2+ transients and modulate cAMP response element binding protein phosphorylation. J. Cell Sci 118, 3131–3140 (2005). [DOI] [PubMed] [Google Scholar]
- 33.Kwong JQ, Henning MS, Starkov AA, Manfredi G, The mitochondrial respiratory chain is a modulator of apoptosis. J. Cell Biol 179, 1163–1177 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Pierro C, Cook SJ, Foets TCF, Bootman MD, Roderick HL, Oncogenic K-Ras suppresses IP3-dependent Ca2+ release through remodelling of the isoform composition of IP3Rs and ER luminal Ca2+ levels in colorectal cancer cell lines. J. Cell Sci 127, 1607–1619 (2014). [DOI] [PubMed] [Google Scholar]
- 35.Shibao K, Fiedler MJ, Nagata J, Minagawa N, Hirata K, Nakayama Y, Iwakiri Y, Nathanson MH, Yamaguchi K, The type III inositol 1,4,5-trisphosphate receptor is associated with aggressiveness of colorectal carcinoma. Cell Calcium 48, 315–323 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Rana M, de Coo I, Diaz F, Smeets H, Moraes CT, An out-of-frame cytochrome B gene deletion from a patient with parkinsonism is associated with impaired complex III assembly and an increase in free radical production. Ann. Neurol 48, 774–781 (2000). [PubMed] [Google Scholar]
- 37.Ryter SW, Kim HP, Hoetzel A, Park JW, Nakahira K, Wang X, Choi AMK, Mechanisms of cell death in oxidative stress. Antioxid. Redox Signal 9, 49–89 (2007). [DOI] [PubMed] [Google Scholar]
- 38.Bell EL, Chandel NS, Genetics of mitochondrial electron transport chain in regulating oxygen sensing. Methods Enzymol. 435, 447–461 (2007). [DOI] [PubMed] [Google Scholar]
- 39.Wise DR, Ward PS, Shay JES, Cross JR, Gruber JJ, Sachdeva UM, Platt JM, DeMatteo RG, Simon MC, Thompson CB, Hypoxia promotes isocitrate dehydrogenase-dependent carboxylation of α-ketoglutarate to citrate to support cell growth and viability. Proc. Natl. Acad. Sci. U.S.A 108, 19611–19616 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Metallo CM, Gameiro PA, Bell EL, Mattaini KR, Yang J, Hiller K, Jewell CM, Johnson ZR, Irvine DJ, Guarente L, Kelleher JK, Vander Heiden MG, Iliopoulos O, Stephanopoulos G, Reductive glutamine metabolism by IDH1 mediates lipogenesis under hypoxia. Nature 481, 380–384 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Armstrong CT, Anderson JL, Denton RM, Studies on the regulation of the human E1 subunit of the 2-oxoglutarate dehydrogenase complex, including the identification of a novel calcium-binding site. Biochem. J 459, 369–381 (2014). [DOI] [PubMed] [Google Scholar]
- 42.Chance B, Schoener B, Oshino R, Itshak F, Nakase Y, Oxidation-reduction ratio studies of mitochondria in freeze-trapped samples. NADH and flavoprotein fluorescence signals. J. Biol. Chem 254, 4764–4771 (1979). [PubMed] [Google Scholar]
- 43.Chance B, Thorell B, Localization and kinetics of reduced pyridine nucleotide in living cells by microfluorometry. J. Biol. Chem 234, 3044–3050 (1959). [PubMed] [Google Scholar]
- 44.Liu Z, Pouli D, Alonzo CA, Varone A, Karaliota S, Quinn KP, Münger K, Karalis KP, Georgakoudi I, Mapping metabolic changes by noninvasive, multiparametric, high-resolution imaging using endogenous contrast. Sci. Adv 4, eaap9302 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Mongeon R, Venkatachalam V, Yellen G, Cytosolic NADH-NAD+ redox visualized in brain slices by two-photon fluorescence lifetime biosensor imaging. Antioxid. Redox Signal 25, 553–563 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Yi J, Luo J, SIRT1 and p53, effect on cancer, senescence and beyond. Biochim. Biophys. Acta 1804, 1684–1689 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Mauthe M, Orhon I, Rocchi C, Zhou X, Luhr M, Hijlkema K-J, Coppes RP, Engedal N, Mari M, Reggiori F, Chloroquine inhibits autophagic flux by decreasing autophagosome-lysosome fusion. Autophagy 14, 1435–1455 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Mizushima N, Yoshimori T, Levine B, Methods in mammalian autophagy research. Cell 140, 313–326 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Ahumada-Castro U, Silva-Pavez E, Lovy A, Pardo E, Molgό J, Cárdenas C, MTOR-independent autophagy induced by interrupted endoplasmic reticulum-mitochondrial Ca2+ communication: A dead end in cancer cells. Autophagy 15, 358–361 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Tretter L, Adam-Vizi V, Alpha-ketoglutarate dehydrogenase: A target and generator of oxidative stress. Philos. Trans. R. Soc. Lond. B Biol. Sci 360, 2335–2345 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.DeBerardinis RJ, Sayed N, Ditsworth D, Thompson CB, Brick by brick: Metabolism and tumor cell growth. Curr. Opin. Genet. Dev 18, 54–61 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Jia S, Xu X, Zhou S, Chen Y, Ding G, Cao L, Fisetin induces autophagy in pancreatic cancer cells via endoplasmic reticulum stress- and mitochondrial stress-dependent pathways. Cell Death Dis. 10, 142 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Wise DR, Thompson CB, Glutamine addiction: A new therapeutic target in cancer. Trends Biochem. Sci 35, 427–433 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Bunik VI, Mkrtchyan G, Grabarska A, Oppermann H, Daloso D, Araujo WL, Juszczak M, Rzeski W, Bettendorff L, Fernie AR, Meixensberger J, Stepulak A, Gaunitz F, Inhibition of mitochondrial 2-oxoglutarate dehydrogenase impairs viability of cancer cells in a cell-specific metabolism-dependent manner. Oncotarget 7, 26400–26421 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Stuart SD, Schauble A, Gupta S, Kennedy AD, Keppler BR, Bingham PM, Zachar Z, A strategically designed small molecule attacks alpha-ketoglutarate dehydrogenase in tumor cells through a redox process. Cancer Metab. 2, 4 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Atlante S, Visintin A, Marini E, Savoia M, Dianzani C, Giorgis M, Sürün D, Maione F, Schnütgen F, Farsetti A, Zeiher AM, Bertinaria M, Giraudo E, Spallotta F, Cencioni C, Gaetano C, α-Ketoglutarate dehydrogenase inhibition counteracts breast cancer-associated lung metastasis. Cell Death Dis. 9, 756 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Denton RM, Hughes WA, Pyruvate dehydrogenase and the hormonal regulation of fat synthesis in mammalian tissues. Int. J. Biochem 9, 545–552 (1978). [DOI] [PubMed] [Google Scholar]
- 58.Demers-Lamarche J, Guillebaud G, Tlili M, Todkar K, Bélanger N, Grondin M, Nguyen AP, Michel J, Germain M, Loss of mitochondrial function impairs lysosomes. J. Biol. Chem 291, 10263–10276 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Baixauli F, Acín-Pérez R, Villarroya-Beltrí C, Mazzeo C, Nuñez-Andrade N, Gabandé-Rodriguez E, Ledesma MD, Blázquez A, Martin MA, Falcón-Pérez JM, Redondo JM, Enríquez JA, Mittelbrunn M, Mitochondrial respiration controls lysosomal function during inflammatory T cell responses. Cell Metab. 22, 485–498 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Cáceres M, Ortiz L, Recabarren T, Romero A, Colombo A, Leiva-Salcedo E, Varela D, Rivas J, Silva I, Morales D, Campusano C, Almarza O, Simon F, Toledo H, Park K-S, Trimmer JS, Cerda O, TRPM4 Is a novel component of the adhesome required for focal adhesion disassembly, migration and contractility. PLOS ONE 10, e0130540 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Alonzo CA, Karaliota S, Pouli D, Liu Z, Karalis KP, Georgakoudi I, Two-photon excited fluorescence of intrinsic fluorophores enables label-free assessment of adipose tissue function. Sci. Rep 6, 31012 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Urra FA, Muñoz F, Córdova-Delgado M, Ramírez MP, Peña-Ahumada B, Rios M, Cruz P, Ahumada-Castro U, Bustos G, Silva-Pavez E, Pulgar R, Morales D, Varela D, Millas-Vargas JP, Retamal E, Ramírez-Rodríguez O, Pessoa-Mahana H, Pavani M, Ferreira J, Cárdenas C, Araya-Maturana R, FR58P1a; a new uncoupler of OXPHOS that inhibits migration in triple-negative breast cancer cells via Sirt1/AMPK/β1-integrin pathway. Sci. Rep 8, 13190 (2018). [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
Fig. S1. Mitochondrial parameters and simultaneous knockdown of types 1 and 3 InsP3R induces cell death in OXPHOS-defective 143BΔcytb cells.
Fig. S2. Pseudohypoxic RCC4 cells rely on constitutive InsP3R-mediated Ca2+ signals.
Fig. S3. Inhibition of constitutive InsP3R-mediated Ca2+ transfer to the mitochondria substantially changes the metabolomic profile of 143BΔcytb cell lines.
Fig. S4. OXPHOS-defective cells rely on constitutive InsP3R-mediated Ca2+ signals.
Fig. S5. Inhibition of constitutive InsP3R-mediated Ca2+ release reduces αKGDH activity and activates SIRT1 and AMPK.
Fig. S6. SIRT1 is necessary for the autophagy induced by inhibition of Ca2+ transfer to mitochondria.
Data Availability Statement
All data needed to evaluate the conclusions in the paper are present in the paper or the Supplementary Materials.