OxPhos defects cause hypermetabolism and reduce lifespan in cells and in patients with mitochondrial diseases

Gabriel Sturm, Kalpita R Karan, Anna S Monzel, Balaji Santhanam, Tanja Taivassalo, Céline Bris, Sarah A Ware, Marissa Cross, Atif Towheed, Albert Higgins-Chen, Meagan J McManus, Andres Cardenas, Jue Lin, Elissa S Epel, Shamima Rahman, John Vissing, Bruno Grassi, Morgan Levine, Steve Horvath, Ronald G Haller, Guy Lenaers, Douglas C Wallace, Marie-Pierre St-Onge, Saeed Tavazoie, Vincent Procaccio, Brett A Kaufman, Erin L Seifert, Michio Hirano, Martin Picard, Gabriel Sturm, Kalpita R Karan, Anna S Monzel, Balaji Santhanam, Tanja Taivassalo, Céline Bris, Sarah A Ware, Marissa Cross, Atif Towheed, Albert Higgins-Chen, Meagan J McManus, Andres Cardenas, Jue Lin, Elissa S Epel, Shamima Rahman, John Vissing, Bruno Grassi, Morgan Levine, Steve Horvath, Ronald G Haller, Guy Lenaers, Douglas C Wallace, Marie-Pierre St-Onge, Saeed Tavazoie, Vincent Procaccio, Brett A Kaufman, Erin L Seifert, Michio Hirano, Martin Picard

Abstract

Patients with primary mitochondrial oxidative phosphorylation (OxPhos) defects present with fatigue and multi-system disorders, are often lean, and die prematurely, but the mechanistic basis for this clinical picture remains unclear. By integrating data from 17 cohorts of patients with mitochondrial diseases (n = 690) we find evidence that these disorders increase resting energy expenditure, a state termed hypermetabolism. We examine this phenomenon longitudinally in patient-derived fibroblasts from multiple donors. Genetically or pharmacologically disrupting OxPhos approximately doubles cellular energy expenditure. This cell-autonomous state of hypermetabolism occurs despite near-normal OxPhos coupling efficiency, excluding uncoupling as a general mechanism. Instead, hypermetabolism is associated with mitochondrial DNA instability, activation of the integrated stress response (ISR), and increased extracellular secretion of age-related cytokines and metabokines including GDF15. In parallel, OxPhos defects accelerate telomere erosion and epigenetic aging per cell division, consistent with evidence that excess energy expenditure accelerates biological aging. To explore potential mechanisms for these effects, we generate a longitudinal RNASeq and DNA methylation resource dataset, which reveals conserved, energetically demanding, genome-wide recalibrations. Taken together, these findings highlight the need to understand how OxPhos defects influence the energetic cost of living, and the link between hypermetabolism and aging in cells and patients with mitochondrial diseases.

Conflict of interest statement

The authors declare no competing interests.

© 2023. The Author(s).

Figures

Fig. 1. Meta-analysis of human studies reveals…
Fig. 1. Meta-analysis of human studies reveals increased energy expenditure and shortened lifespan in primary mitochondrial diseases.
a Overall conceptual model linking mtDNA- and nDNA-related OxPhos defects to impaired metabolic efficiency at the cellular level, impacting whole-body resting energy expenditure and clinical outcomes. b Skeletal muscle biopsy with individual muscle fibers stained with cytochrome c oxidase/succinate dehydrogenase (COX/SDH) histochemistry to reveal functional (brown) and respiratory chain deficient (blue) mitochondria. In the affected cell (middle), three sub-regions showing low, intermediate, and high mtDNA mutation load were captured by laser capture microdissection and subjected to quantitative PCR analysis as in ref. . Subcellular regions with high mtDNA mutation load show elevated mtDNA density, which is predicted to increase the energetic cost due to mitochondrial biogenesis and turnover processes. WT, wild type. c Meta-analysis of human mitochondrial disease cohorts showing elevated resting heart rate (n = 104 controls, 111 patients), d catecholamines (urinary-Cohort 3 and blood-Cohort 6) at rest or during fixed-intensity exercise (n = 38 controls, 19 patients), e whole-body oxygen consumption measured by indirect calorimetry at rest or during the response to mild exercise challenge; one before training, two after training. Slope refers to the rate of increase in VO2 relative to work rate, where a higher slope indicates increased energetic cost for a given work rate (n = 56 controls, 78 patients). f Body mass index (BMI) across mitochondrial disease cohorts and compared to relevant national averages (USA, UK, and Italy combined) (n = 285 controls, 174 patients). g Average life expectancy in individuals with mitochondrial diseases relative to the national average (n = 301 patients). Data are means ± SEM, with % difference between mitochondrial disease and control group were available. h Mortality (age at death) over 10 years (2010–2020) in Cohort 17 compared to national averages for women and men (n = 109 patients). See Table 1 for cohort details. Total n = 225 healthy controls, 690 patients. Only aggregate group means (with or without a measure of variance) were available for some cohorts, so individual participant data is not shown. Standardized effect sizes are quantified as Hedges’ g (g). Overall group comparisons were performed by paired t tests (c and f) or one-sample t tests (d and e), *p < 0.05, **p < 0.01, ****p < 0.0001.
Fig. 2. SURF1 defects decrease metabolic efficiency…
Fig. 2. SURF1 defects decrease metabolic efficiency and cause hypermetabolism without affecting coupling efficiency.
a Schematic of the study design with primary human fibroblasts, coupled with repeated, longitudinal measures of cellular, bioenergetic, and molecular profiling across the lifespan. Three Control and three SURF1 donors (one female, two males in each group) were used for all experiments. b Example oxygen consumption rate (OCR) and extracellular acidification rate (ECAR) obtained from Seahorse measurements of Control and SURF1 cells. c Comparison of average OCR and ECAR values across the cellular lifespan. The specificity of the ECAR signal for glycolysis was verified (see Methods for details). d Lifespan trajectories of ATP production rates (JATP) derived from glycolysis (JATP-Glyc), oxidative phosphorylation (JATP-OxPhos), and total ATP (JATP-Total: Glycolytic- + OxPhos-derived rates) over up to 150 days. Percentages show the average difference between SURF1 and Control across the lifespan. e Lifespan average energy expenditure (EE) by cell line and f corrected for cell volume. g Balance of JATP derived from OxPhos and glycolysis and h quantified SURF1-induced metabolic shift. Dotted lines in (h) denote the range in control cells. i Lifespan trajectory of mtDNAcn and average mtDNAcn at the first 3 time points (early life, days 5–40) and peak value across the lifespan. j Lifespan trajectories and averages of proton leak and k coupling efficiency estimated from extracellular flux measurements of ATP-coupled and uncoupled respiration. n = 3 individuals per group, 7–9 timepoints per individual. Data are means ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001, mixed effects model (fixed effect of control/SURF1 group and days grown, random effects of the donor or cell line).
Fig. 3. Pharmacological inhibition of the mitochondrial…
Fig. 3. Pharmacological inhibition of the mitochondrial F0F1 ATP synthase triggers hypermetabolism.
a Schematic of the study design for fibroblast profiling across the lifespan from three Control donors treated with 1 nM Oligomycin (Oligo). b Lifespan trajectories of JATP (Glycolytic + OxPhos) derived from oxygen consumption rate (OCR) and extracellular acidification rate (ECAR) obtained from Seahorse measurements across the cells’ lifespan (up to 150 days). Percentages show the total average difference between Oligo and Control. c Balance of JATP derived from OxPhos and glycolysis across the lifespan and d oligo-induced metabolic shift. Dotted lines denote the range in control cells. e Relative average lifespan energy expenditure by cell line normalized to control, f corrected for cell volume. g Average of proton leak and h coupling efficiency measures on the Seahorse normalized to control. i Lifespan trajectories and j average mtDNA copy number at the first three time points (early life) and peak value across the lifespan. n = 3 individuals per group, 7–9 timepoints per individual. Data are means ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001, mixed effects model for Oligo vs. control.
Fig. 4. Longitudinal mtDNA deletion profiles in…
Fig. 4. Longitudinal mtDNA deletion profiles in OxPhos deficient SURF1 and Oligo cells.
a RNAseq gene expression results for all MitoCarta 3.0 pathways, plus all mtDNA genes, and the integrated stress response (ISR: average of ATF4, ATF5, CHOP, GDF15). Values for each pathway are computed from the average expression levels of all genes in each pathway, expressed as the median-centered value relative to the youngest control timepoints for each pathway (rows). Each column represents a single timepoints (n = 3–8) along the lifespan of each donor or treatment condition (n = 3 donors, 3 groups). b Gene expression time course of selected mitochondrial pathways from E, expressed on a Log2 scale relative to the first control timepoint (baseline). c 10 kb long-range PCR product resolved by agarose gel electrophoresis for control fibroblasts cultured up to 166 days (P3 to 31), and passage-matched SURF1 and Oligo-treated cells. d Results from mtDNA sequencing and Eklipse analysis. Each line in the circos plots depicts a deletion burden in control (Donor2) and SURF1 (Patient2) and Oligo-treated (Donor2) cells from two (early and mid-lifespan) representative passages. e Timecourse of the number of unique mtDNA deletions in control, SURF1, and Oligo-treated cells. Deletion counts were estimated with a variant call cutoff of >5% heteroplasmy. f Total deletion burden in cells across 150 days of lifespan. Data are mean ± SEM. **P < 0.01, ***P < 0.001, mixed effects model (fixed effect of Control/SURF1/Oligo group and days grown, random effects of donor or cell line).
Fig. 5. OxPhos defects trigger hypersecretion of…
Fig. 5. OxPhos defects trigger hypersecretion of metabokines and age-related cytokines.
a Cytokine dynamics across the lifespan measured on two multiplexes (Luminex) arrays. Cytokine levels are normalized to the number of cells at the time of sampling, shown as Log2 median-centered for each cytokine; samples with undetectable values are shown as gray cells. Columns represent repeated measures (n = 6–8) along the lifespan of each control and SURF1 donor (n = 3 per group). b Comparison of maximum cytokine concentration reached in each of the SURF1 and healthy control donors, showing general upregulation of most metabokines and cytokines. The value for TGF-α is heavily influenced by a single very high value in Donor 3. c Cell-free media GDF15 concentration time course as measured on the Cytokine array. Inset compares early release between 20 and 80 days. d Media IL-6 levels across the cellular lifespan by enzyme-linked immunosorbent assay (ELISA), normalized to the number of cells at the time of sampling. e Media GDF15 levels across the cellular lifespan measured by ELISA, normalized to the number of cells at the time of sampling. Samples with non-detectable values (N.D.) are shown as zero values. f Cell-free mitochondrial DNA (cf-mtDNA) dynamics across the cellular lifespan using qPCR, normalized to the number of cells at the time of sampling. n = 3 per group, 6–13 timepoints per condition. Data are means ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001, mixed effects model (fixed effect of Control/SURF1/Oligo group and days grown, random effects of donor or cell line). Abbreviations: CCL7 C-C motif chemokine ligand 7, IL-8 interleukin 8, CHI3L1 chitinase-3-like protein 1, MMP7 Matrix metallopeptidase 7, IL-6 Interleukin 6, IGFBP-rp1 Insulin-like growth factor binding protein 7, TNF-RII tumor necrosis factor receptor superfamily member 1B, TGF-α tumor growth factor alpha, IFN-γ interferon-gamma, TNFRSF9 TNF receptor superfamily member 9, GDF-15 growth differentiation factor 15, TNF-β tumor necrosis factor beta, Fas Fas cell surface death receptor, CCL3 C-C motif chemokine ligand 7, FSTL1 Follistatin like 1, CCL23 C-C motif chemokine ligand 23, TIMP-1 tissue inhibitor of metallopeptidase 1, CD163 CD163 antigen, Lumican keratan sulfate proteoglycan Lumican, IL-18 interleukin-18, CXCL16 C-X-C motif chemokine ligand 16, Fetuin A alpha 2-HS glycoprotein, ALCAM activated leukocyte cell adhesion molecule, TNF-RI TNF receptor superfamily member 1A, PCSK9 proprotein convertase subtilisin/kexin type 9, TFPI tissue factor pathway inhibitor.
Fig. 6. Mitochondrial defects trigger conserved transcriptional…
Fig. 6. Mitochondrial defects trigger conserved transcriptional remodeling.
a t-distributed stochastic neighbor embedding (t-SNE) of RNAseq data from control, SURF1, and Oligo-treated human fibroblasts across the lifespan. b Overlap of significantly upregulated (red) or downregulated (blue) genes in SURF1 and Oligo groups relative to control (linear mixed effects model, FDR-corrected p value < 0.05). Note, outer group counts include shared counts in the overlapping rings. Gray indicates the diverging direction of regulation between SURF1 and Oligo DEGs. c Expression levels of the top 100 differentially expressed genes in SURF1 (<75 days grown) and Oligo-treated cells (days 35–110). d iPAGE analysis of RNAseq data showing the top 40 enriched gene ontology pathways in top overlapping upregulated and downregulated genes, conserved across both SURF1 and Oligo groups relative to control. Note, −log(p value) > 8 are mapped as dark orange. e Gene expression timecourses of select genes related to the ISR, senescence, nucleotide metabolism, and telomere maintenance. Log2 expression values (TPM) are normalized to the median of the control youngest timepoints. n = 3 donors per group, 3–8 timepoints per donor.
Fig. 7. Mitochondrial defects trigger conserved epigenetic…
Fig. 7. Mitochondrial defects trigger conserved epigenetic remodeling.
a t-SNE of the nuclear DNA methylome from control, SURF1-disease (<75 days grown), and Oligomycin-treated (35–110 days grown) fibroblasts across the replicative lifespan. b Overlap of differentially methylated CpGs (DMPs, left panel) and differentially methylated regions (DMRs, right-panel) generated from mixed effects models. Note, outer group counts include shared counts in the overlapping rings. c Heatmap of top 100 DMPs in SURF1 and Oligo-treated cells. DMPs were ordered by mean methylation difference between groups. d Timecourse of top three hyper- and hypo-DMPs for SURF1 and Oligo-treated groups. e Gene regional map of top hyper- and hypo-DMRs for SURF1-disease and Oligo-treated fibroblasts. 5′–>3′ direction. f Heatmap of top 20 enriched gene ontology pathways in top 1000 hyper- and hypo-DMPs and DMRs overlapping between SURF1 and Oligo-treated groups. Note, −log(P values) > 10 are mapped as dark orange. n = 3 donors per group, 5–11 timepoints per donor/treatment.
Fig. 8. Mitochondrial OxPhos defects decrease lifespan…
Fig. 8. Mitochondrial OxPhos defects decrease lifespan and accelerate telomere shortening.
a Growth curves of control, SURF1, and Oligo-treated cells. Population doublings were determined from both live and dead cell cells at each passage. b Hayflick limit defined as the total number of population doublings achieved before division rate <0.01 divisions/day for at least two passages. c Telomere length per population doubling, d rate of telomere attrition per division, and e terminal telomere length. f Rate of epigenetic aging for control, SURF1, and Oligo-treated cells, calculated from the linear rate between days 25 and 75 (3–4 timepoints/cell line). g Average rate of epigenetic aging across all PC-based clocks. Each datapoint represents a different clock. f, g Significance values were calculated using a multiple comparison two-way ANOVA. n = 3 donors per group, 5–15 timepoints per condition for telomere length. In d, data are the slope estimate for the linear regressions in (c). Data are means ± SEM. *P < 0.05, **P < 0.01.
Fig. 9. Conceptual model including potential sources…
Fig. 9. Conceptual model including potential sources of hypermetabolism in cells and patients with mitochondrial diseases.
OxPhos defects trigger mtDNA instability and cell-autonomous stress responses associated with the hypersecretory phenotype, recapitulating findings in plasma of patients with elevated metabokine and cell-free mitochondrial DNA (cf-mtDNA) levels. These responses are linked to the upregulation of multiple energy-dependent transcriptional programs, including the integrated stress response (ISR). We propose that these processes collectively increase energy consumption, leading to hypermetabolism in patient-derived fibroblasts, and physiological hypermetabolism in affected patients. In dividing human fibroblasts, hypermetabolism-causing OxPhos defects curtail lifespan and accelerate canonical cellular senescence and aging markers, namely telomere length, epigenetic aging, as well as secreted and transcriptional markers.

References

    1. Gorman GS, et al. Mitochondrial diseases. Nat. Rev. Dis. Prim. 2016;2:16080. doi: 10.1038/nrdp.2016.80.
    1. Rossignol R, et al. Mitochondrial threshold effects. Biochem J. 2003;370:751–762. doi: 10.1042/bj20021594.
    1. Mick, E. et al. Distinct mitochondrial defects trigger the integrated stress response depending on the metabolic state of the cell. eLife9, e49178 (2020).
    1. Picard M, et al. Progressive increase in mtDNA 3243A>G heteroplasmy causes abrupt transcriptional reprogramming. Proc. Natl Acad. Sci. USA. 2014;111:E4033–E4042. doi: 10.1073/pnas.1414028111.
    1. Quiros PM, et al. Multi-omics analysis identifies ATF4 as a key regulator of the mitochondrial stress response in mammals. J. Cell Biol. 2017;216:2027–2045. doi: 10.1083/jcb.201702058.
    1. Khan NA, et al. mTORC1 regulates mitochondrial integrated stress response and mitochondrial myopathy progression. Cell Metab. 2017;26:419–428.e415. doi: 10.1016/j.cmet.2017.07.007.
    1. Kuhl, I. et al. Transcriptomic and proteomic landscape of mitochondrial dysfunction reveals secondary coenzyme Q deficiency in mammals. eLife6, e30952 (2017).
    1. Kaspar, S. et al. Adaptation to mitochondrial stress requires CHOP-directed tuning of ISR. Sci. Adv.7, eabf0971 (2021).
    1. Forsstrom S, et al. Fibroblast growth factor 21 drives dynamics of local and systemic stress responses in mitochondrial myopathy with mtDNA deletions. Cell Metab. 2019;30:1040–1054.e1047. doi: 10.1016/j.cmet.2019.08.019.
    1. Chung HK, et al. Growth differentiation factor 15 is a myomitokine governing systemic energy homeostasis. J. Cell Biol. 2017;216:149–165. doi: 10.1083/jcb.201607110.
    1. Suomalainen A, Battersby BJ. Mitochondrial diseases: the contribution of organelle stress responses to pathology. Nat. Rev. Mol. Cell Biol. 2018;19:77–92. doi: 10.1038/nrm.2017.66.
    1. Gorman GS, et al. Perceived fatigue is highly prevalent and debilitating in patients with mitochondrial disease. Neuromuscul. Disord. 2015;25:563–566. doi: 10.1016/j.nmd.2015.03.001.
    1. Kaufmann P, et al. Protean phenotypic features of the A3243G mitochondrial DNA mutation. Arch. Neurol. 2009;66:85–91. doi: 10.1001/archneurol.2008.526.
    1. Porcelli S, Marzorati M, Morandi L, Grassi B. Home-based aerobic exercise training improves skeletal muscle oxidative metabolism in patients with metabolic myopathies. J. Appl. Physiol. (1985) 2016;121:699–708. doi: 10.1152/japplphysiol.00885.2015.
    1. Apabhai S, et al. Habitual physical activity in mitochondrial disease. PLoS ONE. 2011;6:e22294. doi: 10.1371/journal.pone.0022294.
    1. Lowell BB, Shulman GI. Mitochondrial dysfunction and type 2 diabetes. Science. 2005;307:384–387. doi: 10.1126/science.1104343.
    1. Hou Y, et al. Appendicular skeletal muscle mass: a more sensitive biomarker of disease severity than BMI in adults with mitochondrial diseases. PLoS ONE. 2019;14:e0219628. doi: 10.1371/journal.pone.0219628.
    1. Makarieva AM, et al. Mean mass-specific metabolic rates are strikingly similar across life’s major domains: evidence for life’s metabolic optimum. Proc. Natl Acad. Sci. USA. 2008;105:16994–16999. doi: 10.1073/pnas.0802148105.
    1. Basan M, et al. Overflow metabolism in Escherichia coli results from efficient proteome allocation. Nature. 2015;528:99–104. doi: 10.1038/nature15765.
    1. Buttgereit F, Brand MD. A hierarchy of ATP-consuming processes in mammalian cells. Biochem. J. 1995;312:163–167. doi: 10.1042/bj3120163.
    1. Kafri M, Metzl-Raz E, Jona G, Barkai N. The cost of protein production. Cell Rep. 2016;14:22–31. doi: 10.1016/j.celrep.2015.12.015.
    1. Vincent AE, Picard M. Multilevel heterogeneity of mitochondrial respiratory chain deficiency. J. Pathol. 2018;246:261–265. doi: 10.1002/path.5146.
    1. Hoitzing H, et al. Energetic costs of cellular and therapeutic control of stochastic mitochondrial DNA populations. PLoS Comput. Biol. 2019;15:e1007023. doi: 10.1371/journal.pcbi.1007023.
    1. Becker, C. et al. CLPP deficiency protects against metabolic syndrome but hinders adaptive thermogenesis. EMBO Rep.19, e45126 (2018).
    1. Morrow RM, et al. Mitochondrial energy deficiency leads to hyperproliferation of skeletal muscle mitochondria and enhanced insulin sensitivity. Proc. Natl Acad. Sci. USA. 2017;114:2705–2710. doi: 10.1073/pnas.1700997114.
    1. Celotto AM, Chiu WK, Van Voorhies W, Palladino MJ. Modes of metabolic compensation during mitochondrial disease using the Drosophila model of ATP6 dysfunction. PLoS ONE. 2011;6:e25823. doi: 10.1371/journal.pone.0025823.
    1. Agostino A, et al. Constitutive knockout of Surf1 is associated with high embryonic lethality, mitochondrial disease and cytochrome c oxidase deficiency in mice. Hum. Mol. Genet. 2003;12:399–413. doi: 10.1093/hmg/ddg038.
    1. Wedatilake Y, et al. SURF1 deficiency: a multi-centre natural history study. Orphanet J. Rare Dis. 2013;8:96. doi: 10.1186/1750-1172-8-96.
    1. Barends M, et al. Causes of death in adults with mitochondrial disease. JIMD Rep. 2016;26:103–113. doi: 10.1007/8904_2015_449.
    1. Kaufmann P, et al. Natural history of MELAS associated with mitochondrial DNA m.3243A>G genotype. Neurology. 2011;77:1965–1971. doi: 10.1212/WNL.0b013e31823a0c7f.
    1. Keshavan N, Rahman S. Natural history of mitochondrial disorders: a systematic review. Essays Biochem. 2018;62:423–442. doi: 10.1042/EBC20170108.
    1. Trifunovic A, et al. Premature ageing in mice expressing defective mitochondrial DNA polymerase. Nature. 2004;429:417–423. doi: 10.1038/nature02517.
    1. Kujoth GC, et al. Mitochondrial DNA mutations, oxidative stress, and apoptosis in mammalian aging. Science. 2005;309:481–484. doi: 10.1126/science.1112125.
    1. Quintana A, Kruse SE, Kapur RP, Sanz E, Palmiter RD. Complex I deficiency due to loss of Ndufs4 in the brain results in progressive encephalopathy resembling Leigh syndrome. Proc. Natl Acad. Sci. USA. 2010;107:10996–11001. doi: 10.1073/pnas.1006214107.
    1. Quadalti C, et al. SURF1 knockout cloned pigs: early onset of a severe lethal phenotype. Biochim. Biophys. Acta Mol. Basis Dis. 2018;1864:2131–2142. doi: 10.1016/j.bbadis.2018.03.021.
    1. Schrack JA, Knuth ND, Simonsick EM, Ferrucci L. “IDEAL” aging is associated with lower resting metabolic rate: the Baltimore Longitudinal Study of Aging. J. Am. Geriatr. Soc. 2014;62:667–672. doi: 10.1111/jgs.12740.
    1. Jumpertz R, et al. Higher energy expenditure in humans predicts natural mortality. J. Clin. Endocrinol. Metab. 2011;96:E972–E976. doi: 10.1210/jc.2010-2944.
    1. Ruggiero C, et al. High basal metabolic rate is a risk factor for mortality: the Baltimore Longitudinal Study of Aging. J. Gerontol. A Biol. Sci. Med. Sci. 2008;63:698–706. doi: 10.1093/gerona/63.7.698.
    1. Tencerova M, et al. Obesity-associated hypermetabolism and accelerated senescence of bone marrow stromal stem cells suggest a potential mechanism for bone fragility. Cell Rep. 2019;27:2050–2062.e2056. doi: 10.1016/j.celrep.2019.04.066.
    1. Molenaar D, van Berlo R, de Ridder D, Teusink B. Shifts in growth strategies reflect tradeoffs in cellular economics. Mol. Syst. Biol. 2009;5:323. doi: 10.1038/msb.2009.82.
    1. Yang, X. et al. Physical bioenergetics: energy fluxes, budgets, and constraints in cells. Proc. Natl Acad. Sci. USA118, e2026786118 (2021).
    1. Pontzer H. Energy constraint as a novel mechanism linking exercise and health. Physiology. 2018;33:384–393. doi: 10.1152/physiol.00027.2018.
    1. Bobba-Alves N, Juster RP, Picard M. The energetic cost of allostatic load. Psychoneuroendocrinology. 2022;146:105951. doi: 10.1016/j.psyneuen.2022.105951.
    1. Takauji Y, et al. Restriction of protein synthesis abolishes senescence features at cellular and organismal levels. Sci. Rep. 2016;6:18722. doi: 10.1038/srep18722.
    1. Takauji Y, En A, Miki K, Ayusawa D, Fujii M. Combinatorial effects of continuous protein synthesis, ERK-signaling, and reactive oxygen species on induction of cellular senescence. Exp. Cell Res. 2016;345:239–246. doi: 10.1016/j.yexcr.2016.06.011.
    1. Moraes CT, et al. Molecular analysis of the muscle pathology associated with mitochondrial DNA deletions. Nat. Genet. 1992;1:359–367. doi: 10.1038/ng0892-359.
    1. Yu-Wai-Man P, et al. OPA1 mutations cause cytochrome c oxidase deficiency due to loss of wild-type mtDNA molecules. Hum. Mol. Genet. 2010;19:3043–3052. doi: 10.1093/hmg/ddq209.
    1. Lehmann D, et al. Understanding mitochondrial DNA maintenance disorders at the single muscle fibre level. Nucleic Acids Res. 2019;47:7430–7443. doi: 10.1093/nar/gkz472.
    1. Vincent AE, et al. Subcellular origin of mitochondrial DNA deletions in human skeletal muscle. Ann. Neurol. 2018;84:289–301. doi: 10.1002/ana.25288.
    1. Green JA. The heart rate method for estimating metabolic rate: review and recommendations. Comp. Biochem Physiol. A Mol. Integr. Physiol. 2011;158:287–304. doi: 10.1016/j.cbpa.2010.09.011.
    1. Vissing J, Galbo H, Haller RG. Exercise fuel mobilization in mitochondrial myopathy: a metabolic dilemma. Ann. Neurol. 1996;40:655–662. doi: 10.1002/ana.410400416.
    1. Ratheiser KM, Brillon DJ, Campbell RG, Matthews DE. Epinephrine produces a prolonged elevation in metabolic rate in humans. Am. J. Clin. Nutr. 1998;68:1046–1052. doi: 10.1093/ajcn/68.5.1046.
    1. Mehta NM, et al. Accuracy of a simplified equation for energy expenditure based on bedside volumetric carbon dioxide elimination measurement-a two-center study. Clin. Nutr. 2015;34:151–155. doi: 10.1016/j.clnu.2014.02.008.
    1. Taivassalo T, et al. The spectrum of exercise tolerance in mitochondrial myopathies: a study of 40 patients. Brain. 2003;126:413–423. doi: 10.1093/brain/awg028.
    1. Tiranti V, et al. Mutations of SURF-1 in Leigh disease associated with cytochrome c oxidase deficiency. Am. J. Hum. Genet. 1998;63:1609–1621. doi: 10.1086/302150.
    1. Hayflick L, Moorhead PS. The serial cultivation of human diploid cell strains. Exp. Cell Res. 1961;25:585–621. doi: 10.1016/0014-4827(61)90192-6.
    1. Sturm G, et al. Human aging DNA methylation signatures are conserved but accelerated in cultured fibroblasts. Epigenetics. 2019;14:961–976. doi: 10.1080/15592294.2019.1626651.
    1. Brydges, C. R. Effect size guidelines, sample size calculations, and statistical power in gerontology. Innov. Aging3, igz036 (2019).
    1. Mookerjee SA, Gerencser AA, Nicholls DG, Brand MD. Quantifying intracellular rates of glycolytic and oxidative ATP production and consumption using extracellular flux measurements. J. Biol. Chem. 2017;292:7189–7207. doi: 10.1074/jbc.M116.774471.
    1. Schmidt CA, Fisher-Wellman KH, Neufer PD. From OCR and ECAR to energy: Perspectives on the design and interpretation of bioenergetics studies. J. Biol. Chem. 2021;297:101140. doi: 10.1016/j.jbc.2021.101140.
    1. Parrinello S, et al. Oxygen sensitivity severely limits the replicative lifespan of murine fibroblasts. Nat. Cell Biol. 2003;5:741–747. doi: 10.1038/ncb1024.
    1. Jain IH, et al. Hypoxia as a therapy for mitochondrial disease. Science. 2016;352:54–61. doi: 10.1126/science.aad9642.
    1. Guo X, et al. Mitochondrial stress is relayed to the cytosol by an OMA1-DELE1-HRI pathway. Nature. 2020;579:427–432. doi: 10.1038/s41586-020-2078-2.
    1. Cao Y, Wang H, Ouyang Q, Tu Y. The free energy cost of accurate biochemical oscillations. Nat. Phys. 2015;11:772–778. doi: 10.1038/nphys3412.
    1. Rath S, et al. MitoCarta3.0: an updated mitochondrial proteome now with sub-organelle localization and pathway annotations. Nucleic Acids Res. 2021;49:D1541–D1547. doi: 10.1093/nar/gkaa1011.
    1. Bris, C, et al. Improved detection of mitochondrial DNA instability in mitochondrial genome maintenance disorders. Genet. Med.23, 1769–1778 (2021).
    1. Ye K, Lu J, Ma F, Keinan A, Gu Z. Extensive pathogenicity of mitochondrial heteroplasmy in healthy human individuals. Proc. Natl Acad. Sci. USA. 2014;111:10654–10659. doi: 10.1073/pnas.1403521111.
    1. Klein HU, et al. Characterization of mitochondrial DNA quantity and quality in the human aged and Alzheimer’s disease brain. Mol. Neurodegener. 2021;16:75. doi: 10.1186/s13024-021-00495-8.
    1. Tanaka T, et al. Plasma proteomic signature of age in healthy humans. Aging Cell. 2018;17:e12799. doi: 10.1111/acel.12799.
    1. Basisty N, et al. A proteomic atlas of senescence-associated secretomes for aging biomarker development. PLoS Biol. 2020;18:e3000599. doi: 10.1371/journal.pbio.3000599.
    1. Lehtonen, J. M. et al. FGF21 is a biomarker for mitochondrial translation and mtDNA maintenance disorders. Neurology87, 2290-2299 (2016).
    1. Sharma, R. et al. Circulating markers of NADH-reductive stress correlate with mitochondrial disease severity. J. Clin. Invest.131, e136055 (2021).
    1. Lehallier B, et al. Undulating changes in human plasma proteome profiles across the lifespan. Nat. Med. 2019;25:1843–1850. doi: 10.1038/s41591-019-0673-2.
    1. Pinti M, et al. Circulating mitochondrial DNA increases with age and is a familiar trait: Implications for “inflamm-aging”. Eur. J. Immunol. 2014;44:1552–1562. doi: 10.1002/eji.201343921.
    1. Maresca A, et al. Expanding and validating the biomarkers for mitochondrial diseases. J. Mol. Med. 2020;98:1467–1478. doi: 10.1007/s00109-020-01967-y.
    1. Gutierrez JM, et al. Genome-scale reconstructions of the mammalian secretory pathway predict metabolic costs and limitations of protein secretion. Nat. Commun. 2020;11:68. doi: 10.1038/s41467-019-13867-y.
    1. Goodarzi H, Elemento O, Tavazoie S. Revealing global regulatory perturbations across human cancers. Mol. Cell. 2009;36:900–911. doi: 10.1016/j.molcel.2009.11.016.
    1. Hämäläinen RH, et al. Defects in mtDNA replication challenge nuclear genome stability through nucleotide depletion and provide a unifying mechanism for mouse progerias. Nat. Metab. 2019;1:958–965. doi: 10.1038/s42255-019-0120-1.
    1. Wiley CD, et al. Mitochondrial dysfunction induces senescence with a distinct secretory phenotype. Cell Metab. 2016;23:303–314. doi: 10.1016/j.cmet.2015.11.011.
    1. Lozoya OA, et al. Mitochondrial nicotinamide adenine dinucleotide reduced (NADH) oxidation links the tricarboxylic acid (TCA) cycle with methionine metabolism and nuclear DNA methylation. PLoS Biol. 2018;16:e2005707. doi: 10.1371/journal.pbio.2005707.
    1. Lozoya OA, et al. Single nucleotide resolution analysis reveals pervasive, long-lasting DNA methylation changes by developmental exposure to a mitochondrial toxicant. Cell Rep. 2020;32:108131. doi: 10.1016/j.celrep.2020.108131.
    1. Chen DP, Lin YC, Fann CS. Methods for identifying differentially methylated regions for sequence- and array-based data. Brief. Funct. Genomics. 2016;15:485–490.
    1. Oexle K, Zwirner A. Advanced telomere shortening in respiratory chain disorders. Hum. Mol. Genet. 1997;6:905–908. doi: 10.1093/hmg/6.6.905.
    1. Qian W, et al. Chemoptogenetic damage to mitochondria causes rapid telomere dysfunction. Proc. Natl Acad. Sci. USA. 2019;116:18435–18444. doi: 10.1073/pnas.1910574116.
    1. Horvath S, Raj K. DNA methylation-based biomarkers and the epigenetic clock theory of ageing. Nat. Rev. Genet. 2018;19:371–384. doi: 10.1038/s41576-018-0004-3.
    1. Higgins-Chen, A. T. et al. A computational solution for bolstering reliability of epigenetic clocks: Implications for clinical trials and longitudinal tracking. Nat Aging. 2, 644–661 (2022).
    1. Spendiff S, et al. Mitochondrial DNA deletions in muscle satellite cells: implications for therapies. Hum. Mol. Genet. 2013;22:4739–4747. doi: 10.1093/hmg/ddt327.
    1. Latorre-Pellicer A, et al. Mitochondrial and nuclear DNA matching shapes metabolism and healthy ageing. Nature. 2016;535:561–565. doi: 10.1038/nature18618.
    1. Caporali L, et al. Peculiar combinations of individually non-pathogenic missense mitochondrial DNA variants cause low penetrance Leber’s hereditary optic neuropathy. PLoS Genet. 2018;14:e1007210. doi: 10.1371/journal.pgen.1007210.
    1. Raule N, et al. The co-occurrence of mtDNA mutations on different oxidative phosphorylation subunits, not detected by haplogroup analysis, affects human longevity and is population specific. Aging Cell. 2014;13:401–407. doi: 10.1111/acel.12186.
    1. Lan G, Sartori P, Neumann S, Sourjik V, Tu Y. The energy-speed-accuracy tradeoff in sensory adaptation. Nat. Phys. 2012;8:422–428. doi: 10.1038/nphys2276.
    1. Inak G, et al. Defective metabolic programming impairs early neuronal morphogenesis in neural cultures and an organoid model of Leigh syndrome. Nat. Commun. 2021;12:1929. doi: 10.1038/s41467-021-22117-z.
    1. Jeppesen TD, Orngreen MC, van Hall G, Haller RG, Vissing J. Fat metabolism during exercise in patients with mitochondrial disease. Arch. Neurol. 2009;66:365–370. doi: 10.1001/archneurol.2009.24.
    1. Saleem A, et al. Polymerase gamma mutator mice rely on increased glycolytic flux for energy production. Mitochondrion. 2015;21:19–26. doi: 10.1016/j.mito.2014.12.001.
    1. Picard M, et al. Mitochondrial functions modulate neuroendocrine, metabolic, inflammatory, and transcriptional responses to acute psychological stress. Proc. Natl Acad. Sci. USA. 2015;112:E6614–E6623. doi: 10.1073/pnas.1515733112.
    1. McManus MJ, et al. Mitochondrial DNA variation dictates expressivity and progression of nuclear DNA mutations causing cardiomyopathy. Cell Metab. 2019;29:78–90.e75. doi: 10.1016/j.cmet.2018.08.002.
    1. Kirkman MA, et al. Gene-environment interactions in Leber hereditary optic neuropathy. Brain. 2009;132:2317–2326. doi: 10.1093/brain/awp158.
    1. Rabenstein A, et al. Smoking and alcohol, health-related quality of life and psychiatric comorbidities in Leber’s Hereditary Optic Neuropathy mutation carriers: a prospective cohort study. Orphanet. J. Rare Dis. 2021;16:127. doi: 10.1186/s13023-021-01724-5.
    1. Guimier, A. et al. PPA2-associated sudden cardiac death: extending the clinical and allelic spectrum in 20 new families. Genet. Med.23, 2415–2425 (2021).
    1. Suter PM, Schutz Y, Jequier E. The effect of ethanol on fat storage in healthy subjects. N. Engl. J. Med. 1992;326:983–987. doi: 10.1056/NEJM199204093261503.
    1. Shelmet JJ, et al. Ethanol causes acute inhibition of carbohydrate, fat, and protein oxidation and insulin resistance. J. Clin. Invest. 1988;81:1137–1145. doi: 10.1172/JCI113428.
    1. Karan, K. R. et al. Mitochondrial respiratory capacity modulates LPS-induced inflammatory signatures in human blood. Brain Behav. Immun. Health5, 100080 (2020).
    1. Ganeshan K, et al. Energetic trade-offs and hypometabolic states promote disease tolerance. Cell. 2019;177:399–413.e312. doi: 10.1016/j.cell.2019.01.050.
    1. Ganetzky RD, et al. Congenital hypermetabolism and uncoupled oxidative phosphorylation. N. Engl. J. Med. 2022;387:1395–1403. doi: 10.1056/NEJMoa2202949.
    1. Luft R, Ikkos D, Palmieri G, Ernster L, Afzelius B. A case of severe hypermetabolism of nonthyroid origin with a defect in the maintenance of mitochondrial respiratory control: a correlated clinical, biochemical, and morphological study. J. Clin. Invest. 1962;41:1776–1804. doi: 10.1172/JCI104637.
    1. Bates MG, et al. Defining cardiac adaptations and safety of endurance training in patients with m.3243A>G-related mitochondrial disease. Int. J. Cardiol. 2013;168:3599–3608. doi: 10.1016/j.ijcard.2013.05.062.
    1. Newman J, et al. Preliminary Evaluation of Clinician Rated Outcome Measures in Mitochondrial Disease. J. Neuromuscul. Dis. 2015;2:151–155. doi: 10.3233/JND-140061.
    1. Galna B, et al. Discrete gait characteristics are associated with m.3243A>G and m.8344A>G variants of mitochondrial disease and its pathological consequences. J. Neurol. 2014;261:73–82. doi: 10.1007/s00415-013-7129-2.
    1. Strauss KA, et al. Severity of cardiomyopathy associated with adenine nucleotide translocator-1 deficiency correlates with mtDNA haplogroup. Proc. Natl Acad. Sci. USA. 2013;110:3453–3458. doi: 10.1073/pnas.1300690110.
    1. Delaney NF, et al. Metabolic profiles of exercise in patients with McArdle disease or mitochondrial myopathy. Proc. Natl Acad. Sci. USA. 2017;114:8402–8407. doi: 10.1073/pnas.1703338114.
    1. Jeppesen TD, Orngreen MC, Van Hall G, Vissing J. Lactate metabolism during exercise in patients with mitochondrial myopathy. Neuromuscul. Disord. 2013;23:629–636. doi: 10.1016/j.nmd.2013.05.007.
    1. Heinicke K, et al. Exertional dyspnea in mitochondrial myopathy: clinical features and physiological mechanisms. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2011;301:R873–R884. doi: 10.1152/ajpregu.00001.2011.
    1. Grassi B, et al. Metabolic myopathies: functional evaluation by analysis of oxygen uptake kinetics. Med Sci. Sports Exerc. 2009;41:2120–2127. doi: 10.1249/MSS.0b013e3181aae96b.
    1. Grassi B, et al. Impaired oxygen extraction in metabolic myopathies: detection and quantification by near-infrared spectroscopy. Muscle Nerve. 2007;35:510–520. doi: 10.1002/mus.20708.
    1. Eom S, et al. Cause of death in children with mitochondrial diseases. Pediatr. Neurol. 2017;66:82–88. doi: 10.1016/j.pediatrneurol.2016.10.006.
    1. Pontzer H, et al. Daily energy expenditure through the human life course. Science. 2021;373:808–812. doi: 10.1126/science.abe5017.
    1. Picard M, et al. Mitochondrial dysfunction and lipid accumulation in the human diaphragm during mechanical ventilation. Am. J. Respir. Crit. Care Med. 2012;186:1140–1149. doi: 10.1164/rccm.201206-0982OC.
    1. Krishnan KJ, Bender A, Taylor RW, Turnbull DM. A multiplex real-time PCR method to detect and quantify mitochondrial DNA deletions in individual cells. Anal. Biochem. 2007;370:127–129. doi: 10.1016/j.ab.2007.06.024.
    1. Sturm, G. et al. A Multi-omics and bioenergetics longitudinal aging dataset in primary human fibroblasts with mitochondrial perturbations. Sci. Data. 9, 751 (2022).
    1. Brand, M. D. & Nicholls D. G. Assessing mitochondrial dysfunction in cells. Biochem. J.2, 297–312 (2011).
    1. Goudenege D, et al. eKLIPse: a sensitive tool for the detection and quantification of mitochondrial DNA deletions from next-generation sequencing data. Genet. Med. 2019;21:1407–1416. doi: 10.1038/s41436-018-0350-8.
    1. Ware SA, et al. An automated, high-throughput methodology optimized for quantitative cell-free mitochondrial and nuclear DNA isolation from plasma. J. Biol. Chem. 2020;295:15677–15691. doi: 10.1074/jbc.RA120.015237.
    1. Belmonte FR, et al. Digital PCR methods improve detection sensitivity and measurement precision of low abundance mtDNA deletions. Sci. Rep. 2016;6:25186. doi: 10.1038/srep25186.
    1. Bray NL, Pimentel H, Melsted P, Pachter L. Near-optimal probabilistic RNA-seq quantification. Nat. Biotechnol. 2016;34:525–527. doi: 10.1038/nbt.3519.
    1. Ge SX, Jung D, Yao R. ShinyGO: a graphical gene-set enrichment tool for animals and plants. Bioinformatics. 2020;36:2628–2629. doi: 10.1093/bioinformatics/btz931.
    1. Cawthon RM. Telomere measurement by quantitative PCR. Nucleic Acids Res. 2002;30:e47. doi: 10.1093/nar/30.10.e47.
    1. Lin J, et al. Analyses and comparisons of telomerase activity and telomere length in human T and B cells: insights for epidemiology of telomere maintenance. J. Immunol. Methods. 2010;352:71–80. doi: 10.1016/j.jim.2009.09.012.

Source: PubMed

3
Tilaa