Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Dec 1.
Published in final edited form as: Nat Rev Mol Cell Biol. 2022 Jul 8;23(12):817–835. doi: 10.1038/s41580-022-00506-6

Mechanisms of Mitochondrial Respiratory Adaptation

Christopher F Bennett 1,2,3, Pedro Latorre-Muro 1,2,3, Pere Puigserver 1,2
PMCID: PMC9926497  NIHMSID: NIHMS1871472  PMID: 35804199

Abstract

Mitochondrial energetic adaptations encompass a plethora of conserved processes that maintain cell and organismal fitness and survival in the changing environment by adjusting respiratory capacity of mitochondria. These mitochondrial responses are governed by general principles of regulatory biology exemplified by changes in gene expression, protein translation, protein complex formation, transmembrane transport, enzymatic activities and metabolite levels. These changes can promote mitochondrial biogenesis and membrane dynamics that in turn support mitochondrial respiration. Main regulatory components of mitochondrial energetic adaptation include the transcription coactivator PGC1α and associated transcription factors, mTOR and ER stress signalling, TOM70-dependent mitochondrial protein import, cristae remodelling protein complexes MICOS and OPA1, lipid remodelling, and the assembly and metabolite-dependent regulation of respiratory complexes. These adaptive molecular and structural mechanisms increase respiration to maintain basic processes specific to cell types and tissues. Failure to execute these regulatory responses cause cell damage and inflammation or senescence, compromising cell survival and the ability to adapt to energetically demanding conditions. Thus, mitochondrial adaptive cellular processes are important for physiological responses including to nutrient availability, temperature, and physical activity, and their failure leads to pathologies associated with mitochondrial dysfunction such as metabolic and age-associated diseases and cancer.

Introduction

All biological processes are subject to general regulatory principles that include transcription, translation, post-translational and metabolite-dependent activity that define adaptive (acute or chronic) responses. These dynamic responses in turn are sustained and/or attenuated through feedback or feed forward mechanisms that ultimately dictate cellular health and fitness. Functional adaptations of mitochondria do not escape these principles. Carbohydrates, amino acids, and lipids are imported into the cell and delivered to mitochondria, where biosynthetic reactions occur and ATP is generated using the respiratory chain in a process referred to as oxidative phosphorylation (OXPHOS). Consequently, mechanisms of crosstalk between the mitochondria and nucleus or mitochondria and cytosol evolved to adapt cells, tissues, and organisms to energetically demanding conditions present during physiological or disease-associated responses.

Mitochondrial respiration is supported by the assembly of respiratory complexes(Box 1), and is coupled to the breakdown of metabolic fuels, generation of biosynthetic intermediates, mitochondrial redox, and metabolite import/export1,2. Mitochondrial respiratory complexes I-V are multimeric protein complexes that reside in inner mitochondrial membrane invaginations termed cristae [G]. For complexes I, III, IV, and V, assembly requires the prior synthesis of subunits encoded by both the nuclear and mitochondrial genomes. The proton gradient across the inner mitochondrial membrane generated by the proton pumping activity of complexes I, III, and IV is paired to phosphorylation of ADP to ATP at complex V3. In thermogenic brown and beige adipose tissues [G], complex V is not highly represented and instead, futile cycles of electron transport4,5 dissipate the proton gradient to promote thermogenesis rather than ATP production6,7. In this regard, brown adipose tissue mitochondria are very unique and also exhibit extremely high mitochondrial content and cristae density, and are characterized by expression of uncoupling protein 1 [G]6. A byproduct of normal respiratory metabolism exacerbated under certain stimuli such as mitochondrial dysfunction, ischemia-reperfusion events8,9, and cold exposure in brown adipose tissue10, among others, are reactive oxygen species (ROS). These species contribute to signalling events from the mitochondria to the cytosol or nucleus but, if unchecked, may also damage cellular structures including DNA. Aside from their catabolic functions, mitochondria are central regulators of cell death11, govern cellular Ca2+ levels through sequestration and release12, and synthesize certain molecules such as fatty acids, amino acids, nucleotides, haem [G], and iron-sulfur clusters [G]13. Mitochondria are therefore fundamental to cellular metabolism and physiology, and not surprisingly, are downstream targets for cellular responses and adaptations that control their activities.

Box 1 |. Mitochondrial respiratory chain and bioenergetics.

The respiratory chain is composed of five multisubunit complexes that ultimately generate ATP, the energy equivalent of the cell. These complexes are defined as complex I (NADH-UQ oxidoreductase), complex II (succinate dehydrogenase), complex III (UQH2–cyt c oxidoreductase), complex IV (cytochrome c oxidase), and complex V (ATP synthase). In addition to electron transport function, complexes I, III, and IV are oxidation-reduction proton pumps that produce the proton gradient across the mitochondrial inner membrane. The mobile electron carriers in the electron transport chain are ubiquinone/ubiquinol (UQ/UQH2) also known as Coenzyme Q10 (CoQ10) that diffuse within the inner membrane and cytochrome c that is located on the intermembrane space side of the inner membrane. UQH2 donates electrons to complex III, while cytochrome c donates to complex IV.

From complexes I-IV, electrons are transferred from a redox potential span of 1.1V from the NAD+/NADH redox couple to the O2/2H2O couple257. NADH with a midpoint potential of −320 mV, serves as electron shuttle between the matrix dehydrogenases of the TCA cycle and complex I of the respiratory chain257. Many other enzymes with a more positive redox potential cannot thermodynamically donate electrons to NADH and instead, donate electrons directly to the UQ/UQH2 pool, bypassing complex I257. These include the flavoproteins succinate dehydrogenase, glycerol-3-phosphate dehydrogenase, dihydroorotate dehydrogenase (pyrimidine synthesis), and ETF–ubiquinone oxidoreductase (fatty acid oxidation). The redox components within the respiratory chain are flavoproteins, cytochromes containing haem, iron–sulfur (Fe-S) proteins, ubiquinone, and copper that is tightly bound to certain protein subunits. Electron transport paths across the respiratory complexes are reviewed elsewhere257259, however, recent cryo-EM structural information on respiratory complexes and supercomplexes are enabling us to address unresolved and novel questions on the structure-function relations that drive mitochondrial respiratory activity105,178,250,260262.

The rates of electron transport are regulated by multiple mechanisms in cells. Downstream targets include i) allosteric regulation of respiratory enzymes, ii) enzyme content (e.g. respiratory complex abundance), iii) cofactor levels (e.g. NAD+/NADH, FAD/FADH2, CoQ/CoQH2, cytochrome c) mitochondrial density or membrane morphology such as cristae folding. Allosteric control of respiratory enzymes occurs by NAD+/NADH, ATP and AMP, CoA/acetyl CoA, and Ca2+. For example, NADH is an allosteric inhibitor of TCA dehydrogenases263 and ATP is of pyruvate dehydrogenase, isocitrate dehydrogenase263, and cytochrome-c oxidase/complex IV264. Thus, when a cell enters a high energy state defined by high NADH and ATP levels, TCA and respiratory chain activity is attenuated. By contrast, when a cell requires ATP defined by high ADP/ATP ratio or increased AMP levels, TCA and respiratory chain activity are induced.

Cristae provide proper spatial distribution to the respiratory chain complexes1. These structures originate from inner mitochondrial membrane protuberances and elongate inside the mitochondrial matrix14, which is mediated by a multi-subunit complex in the inner mitochondrial membrane known as the mitochondrial contact site and cristae organizing system (MICOS)14. Mitochondrial dynamin like GTPase OPA1 (refs.15,16) and phospholipids17,18 transported from other organelles or synthesized within mitochondria also participate in cristae biogenesis. The dynamic remodelling of cristae architecture19 adapts to cellular metabolic requirements and synchronizes with the respiratory chain to increase the respiratory capacity of cells20,21. The control of mitochondrial cristae dynamics is exerted by the endoplasmic reticulum (ER), which senses cellular metabolic and energetic fluctuations and triggers downstream stress responses21. Physical contacts known as mitochondrial–ER contacts (MERCs) create a unique environment of protein–protein interactions to regulate cristae formation in addition to other essential processes of mitochondrial dynamics such as fission, fusion or mitophagy22. ER stress [G] promotes cristae formation through processes that coordinate mitochondrial protein import21, lipid synthesis18, and mitochondrial respiratory complex assembly20 to sustain mitochondrial function, cell adaptation and survival.

Cold exposure, dietary nutrients, or physical activity are sensed at the organismal and cellular level activating hormonal, metabolite, or protein signalling that control mitochondrial-linked gene transcription and mRNA translation to enhance respiratory capacity (Fig. 1). This is coupled with the assembly and activity of new, functional respiratory complexes within mitochondrial cristae. In some cases, such as cold exposure in brown adipocytes23,24 or exercise in skeletal muscle25,26, this adaptation is linked with mitochondrial biogenesis that involves the synthesis and incorporation of new mtDNA, protein, and membrane into pre-existing mitochondria, thereby increasing mitochondrial mass. Mechanisms of cross-compartment translational control also exist to synchronize synthesis of nuclear-encoded respiratory chain subunits with mitochondrial-encoded subunits to ensure proper stoichiometry of respiratory complexes and preserve mitochondrial protein homeostasis2731. Membrane synthesis, membrane dynamics (fission and fusion) further contribute to the organization of mitochondrial membranes and the embedded protein complexes.

Figure 1 |. Regulation of mitochondrial respiratory capacity.

Figure 1 |

Cells and tissues experience different external and internal conditions that require mitochondrial adaptation to support ATP generation or heat production. Most cells utilize the mitochondrial respiratory chain for ATP generation, whereby ATP synthesis at respiratory complex V (CV, ATP synthase) is coupled to the electron transport through the respiratory complexes I-IV (CI, CII, CIII2, CIV); however, certain cells such as brown and beige adipocytes produce heat through oxidation of metabolic substrates paired with uncoupled respiration. Multiple levels of regulation of mitochondrial respiration: from transcriptional and translation control to protein (post-translational) control. These processes increase mitochondrial biogenesis — depicted by an increase in mitochondrial number — and respiratory capacity. In the latter case, cellular pathways influence precursor protein import, membrane/cristae dynamics, respiratory chain assembly and superassembly, and phospholipid composition or remodelling to either increase the concentration of respiratory complexes in the inner mitochondrial membrane and/or their activity. Protein structures used in this figure are sourced from PDB: 5gup250 (CI+CIII2+CIV), 1zoy251 (CII), 6zpo252 (CV), 1okc253 (AAC), 2lck254 (UCP2). AAC, ADP/ATP carrier; CKMT, mitochondrial creatine kinase; Cr, creatine; P-Cr, creatine phosphate; Pi, inorganic phosphate; Q, ubiquinone; QH2, ubiquinol.

In this Review, we discuss the different regulatory mechanisms that modulate mitochondrial respiration to facilitate cellular adaptation to different metabolic and growth conditions. In particular, we focus on transcriptional, translational, post-translational and metabolite control mechanisms that promote mitochondrial respiratory chain assembly and cristae biogenesis under physiological conditions of high energy demand. Throughout, we highlight the main cellular components that execute these regulatory mechanisms and their effects on cellular and organismal physiology.

Transcriptional regulation mechanisms

Except for a few mitochondrial proteins and RNAs encoded by the mitochondrial DNA, the rest of proteins in this organelle are synthesized from mRNAs produced from the nuclear DNA32. Activation of transcriptional programs supply mitochondrial proteins that increase mitochondrial biogenesis and respiratory capacity33. The most studied transcriptional axis that controls mitochondrial respiratory adaptation in mammals are the gene expression networks centred on PGC1α coactivator, which are active in different energy-demanding conditions such as cold23, physical activity34,35, perinatal cardiac adaptation36, brain dopaminergic activity37,38, or kidney energetic demands39, among others (Fig. 2). How exactly PGC1α is able to sense the different signals required for these various responses is not known. One model is that PGC1α-interacting factors activated by specific stimuli provide context-dependent, cellular adaptations including mitochondrial biogenesis and upregulation of large sets of nuclear-encoded mitochondrial genes involved in specific metabolic, energetic, and structural functions40,41. In addition, PGC1α itself is regulated at the mRNA and protein level (see following section) and by several post-translational modifications such as phosphorylation and acetylation that occur in response to the different signals42.

Figure 2 |. Transcriptional control of mitochondrial biogenesis through PGC1α.

Figure 2 |

PPARγ coactivator-1α (PGC1α) is a major transcriptional regulator of mitochondrial function. It is activated through multiple upstream stimuli including cold exposure and exercise. These signals converge on PGC1α either through the upregulation of PPARGC1A mRNA or stabilization of the protein. Once PGC1α accumulates, it interacts with various transcription factors (TFs), including several nuclear receptors [G] to promote mitochondrial gene expression that together increase mitochondrial functional capacity. The transcriptional network of PGC1α is complex, but one intriguing model is that specificity of gene expression programs is designated through interaction with specific transcription factors to control diverse aspects of mitochondrial function and biogenesis across numerous cell types. Another emerging mechanism of PGC1α-mediated regulation is its direct binding to various mRNAs (including Cluh involved in RNA granules as well as Slc25a25, which encodes a mitochondrial solute carrier (see also. Fig. 3)) through Ser/Arg-rich (RS) domains and RNA recognition motif (RRM). How PGC1α mRNA binding controls mRNA expression, processing, or export is unclear and an exciting area of future discovery. CAMKII, Calcium/calmodulin-dependent protein kinase II; PKA, protein kinase A; AD, activation domain; RD, repression domain.

PGC1α is highly expressed in energy-demanding tissues such as liver, brown adipose tissue, kidney, heart or skeletal muscle23. It associates with transcription complexes linked to different sets of processes regulating mitochondrial function including mitochondrial DNA replication43, transcription23, translation44, OXPHOS24, haem biosynthesis45, iron-sulfur clusters biogenesis33, TCA cycle46, protein import47,48, and metabolite trafficking49. The mechanism of action of PGC1α is through physical interaction with transcription factors (including nuclear respiratory factors (NRF1 and NRF2)24, oestrogen-related receptors (ERRs)50, Yin Yang 1 (YY1)51, and peroxisome proliferator-activated receptors (PPARs)) that bind to promoter regions and activate mitochondrial genes promoting mitochondrial biogenesis (Fig. 2)23. The magnitude of PGC1α-mediated gene activation depends on additional recruitment of chromatin remodelling factors including the p300/CBP acetyl transferase52 or the mediator complex [G]53. Of note, although PGC1α is an important component of stress-induced mitochondrial biogenesis, in some cases, the response might not fully require this coactivator but rather other isoforms or family members such as PGC1β54. In addition, a novel mechanism of PGC1α to regulate gene expression, outside of transcription: PGC1α is able to directly bind to intronic regions of a subset of mRNAs encoding metabolic genes (including mitochondrial ATP transporter Slc25a25) via its RNA binding domains49,55.

In response to different external stimuli, cellular signalling transduction pathways are used to coordinate transcriptional, translational, and post-translational processes that increase mitochondrial biogenesis and respiration. Exposure to lower temperatures in brown/beige fat or physical activity in skeletal muscle exemplify potent signals that stimulate mitochondrial biogenesis and respiratory adaption5658. Cold promotes secretion of catecholamines [G] that activate cAMP signalling to stimulate mitochondrial biogenesis42, which involves the activity of energy-sensing pathways centred on mTOR [G]59 and AMPK [G]60. These signal transduction pathways control, among other factors, transcriptional activity of PGC1α and mitochondrial adaptation in low temperature environments. Another signal downstream of cAMP is the ER stress kinase, PERK, that controls mitochondrial protein import and cristae formation21,61. Cessation of adrenergic signalling and cAMP deactivates the mitochondrial biogenesis program and promotes mitophagy signals to eliminate mitochondrial mass62. Some of the pathways that induce mitochondrial adaptation during exercise in the muscle are similar to those conferred by cold such as AMPK or cAMP6365; another major pathway involves Ca2+ signalling35,66.

Translational regulation

Mitochondria require stringent regulation of cytosolic mRNA translation and subsequent protein import to dynamically respond to energetic and environmental stimuli. There are multiple points of regulation of translation of mitochondrial proteins including translation initiation, targeting of mRNAs to different subcellular locations for protein synthesis, and, for the components of the respiratory chain, synchronized co-translational assembly of nuclear-encoded and mitochondria-encoded subunits into respiratory complexes (Fig. 3). In this section, we describe how these translational control mechanisms communicate with metabolic stimuli to adapt mitochondrial respiration to the needs of the cell.

Figure 3 |. Translational control of mitochondrial respiratory chain assembly.

Figure 3 |

Multiple mechanisms regulate translation of nuclear-encoded mitochondrial mRNAs. mTORC1 is the primary driver of translation through inhibitory phosphorylation of 4E-BP1/2, releasing its interaction from the cap-binding protein eIF4E. This stimulates eIF4F formation on the 5’ cap of mRNAs, recruitment of the 40S ribosome to the mRNA, formation of the AUG-48S initiation complex, and finally the 80S initiation complex preceding translation. Many mRNAs important for energy generation, including mRNAs encoding mitochondrial proteins are highly sensitive to the phosphorylation status of 4E-BPs. mTORC1 also controls energy metabolism by stimulating the activity of several transcriptional regulators such as PPARγ coactivator-1α (PGC1α). Hence mitochondrial energetics is coupled to nutritional status via mTORC1. Nevertheless, certain mRNAs — heavily enriched for mitochondrial mRNAs — contain short 5’ UTRs and in many cases translation initiator elements termed TISU. TISU elements enable efficient translation initiation of short 5’ UTR mRNAs, even when global protein synthesis is impaired from energetic defects. PGC1α is also regulated at the translational level through upstream open reading frame (uORF)-mediated translational repression. If this mechanism occurs under certain physiological conditions or negatively regulated by mTORC1, similar to the uORF-mediated regulation of ER-stress factor ATF4255, is not yet known. Translation of the majority of mRNAs for mitochondrial proteins occurs in the cytosol and resulting peptides are stabilized by heat shock proteins (HSPs) before delivery to the TOM40 channel. For a subset of mRNAs, co-translational import at the outer mitochondrial membrane (OMM) through interaction with TOM70 occurs. Specific mRNAs involved in fasting or cold responses are also stabilized in RNA granules, supporting their translational capacity. One class of RNA granules is formed by CLUH, which associates with mitochondrial mRNAs and also acts to sequester mTORC1 and RNA binding proteins G3BP1/2 to promote mitophagy. Another, termed stress granules, can contain the RNA binding protein FAM195A critical for branched-chain amino acid metabolism and thermogenesis. Paired with the influx of nuclear-encoded proteins, mitochondria adapt their protein synthesis rates via sensor proteins and complexes such the mitochondrial RNA binding protein LRPPRC and mitochondrial translation regulation assembly intermediate of cytochrome c oxidase (MITRAC). These mechanisms coordinate respiratory chain assembly through the assembly of imported peptides with mitochondrial-encoded peptides (all of which encode components of the respiratory chain). IMM, inner mitochondrial membrane.

mTORC1–4E-BP1 regulates translation of nuclear-encoded mitochondrial mRNAs

Many metabolic signals including insulin–PI3K–AKT signalling, ATP sensing through AMPK67,68, and amino acid levels69,70 regulate the nutrient sensor mTORC1, enabling mTORC1 to efficiently integrate metabolic information to control mitochondrial respiratory function. mTORC1 promotes translation of the majority of nuclear-encoded mRNAs through regulation of mRNA 5′ cap-binding protein eIF4E (eukaryotic initiation factor 4E) that recognizes the m7GpppN cap structure at the 5’ end of eukaryotic mRNAs71. mTORC1 phosphorylates eukaryotic initiation factor 4E binding proteins (4E-BP1/2)72 to prevent their inhibitory action towards eIF4E and the assembly of the eIF4F translation initiation complex on mRNAs (reviewed in73). Along with its role in translation, mTORC1 also phosphorylates chromatin modifier YY1 that interacts with PGC1α and targets it to mitochondrial gene promoters47,51, syncing transcriptional and translational mitochondrial biogenesis programs.

Cap-dependent translation is intrinsically linked to intracellular energy availability and mTORC1 activity. During energetic surplus, mTORC1 promotes eIF4E binding to nuclear-encoded mRNAs through phosphorylation of 4E-BP1/2. Although most nuclear-encoded mRNAs require eIF4E for cap-dependent translation, certain mRNA classes defined by their 5’ UTR sequences and structures are particularly sensitive to eIF4E cap binding7478. A class of these mRNAs contain short 5’ UTRs (<30 nucleotides), often with Translation Initiator of Short 5′ UTR (TISU) elements [G] that function both in transcription, through interaction with YY1, and translation [REF]. Nuclear-encoded mitochondrial mRNAs are enriched with TISU elements such as respiratory complex subunits NDUFS4, NDUFS6, UQCRC1, ATP5O, and numerous mitochondrial ribosomal proteins7678. Presence of TISU enables efficient translation initiation under translational stress from glucose deprivation or AMPK agonism7981, which mildly affect 4EBP1/2 phosphorylation. This facilitates mitochondrial ATP production during periods of energetic insufficiency. However, potent inhibition of 4E-BP1/2 phosphorylation with mTORC1 inhibitors impairs translation of mitochondrial mRNAs with short 5’ UTRs (TISU and non-TISU)78 and decreases mitochondrial respiration, TCA metabolism, ATP turnover, mtDNA content, and mitochondrial mass76. Intriguingly, cell survival and mitochondrial network integrity is preserved under these conditions through decreased mTORC1-dependent translation of the mitochondrial fission factor MTFP182. Overall, mTORC1 regulated protein synthesis integrates cellular energy status with dynamic and rapid regulation of mitochondrial respiration and ATP production through selective translation of nuclear-encoded mitochondrial genes.

Translation of PGC1α through alternative open reading frames

An additional layer of translational regulation of mitochondrial function involves repressive upstream open reading frames (uORFs) in the PPARGC1A mRNA (encoding PGC1α)83. uORFs are a major class of genetic element within the 5’ UTR of mRNAs that regulate translation initiation of the main coding sequence. In fact, ~ 50% of human transcripts contain at least one uORF, indicating selective pressure for maintaining these elements84,85. Translational control mediated by uORFs is primarily evident in stress responsive conditions, where the eIF2·GTP-Met-tRNAMet ternary complex necessary for initiation of translation becomes depleted, permitting ribosomal scanning downstream of the inhibitory uORF. PPARGC1A mRNA from humans to flies contain uORF(s) that limits PGC1α protein levels83. Intriguingly, this element and regulatory paradigm is absent from blue fin tuna PPARGC1A sequence83, consistent with its high oxidative muscle fibre and mitochondrial content. Mice containing a PPARGC1A uORFTAA mutation show increased tolerance to acute kidney injury (AKI)83, which is associated with PGC1α suppression86,87. Future research is needed to establish whether uORF translational control of PGC1α is a primary mechanism to promote mitochondrial adaptation under physiological stimuli (e.g. exercise) or stress conditions that lead to reduced translation initiation ternary complex levels.

Role of RNA granules in translation control of mitochondrial transcripts

Other post-transcriptional regulatory mechanisms occur in distinct, membraneless organelles termed RNA granules to adapt mitochondrial respiratory function to metabolic stimuli. Clustered mitochondria homologue (CLUH) is a cytosolic RNA-binding protein that localizes to RNA granules to preserve translational capacity of target mRNAs required under nutrient deprivation that affect respiratory chain function, TCA metabolism, fatty acid oxidation, and amino acid catabolism8890. This is supported by CLUH interaction with ribosomal proteins, translation factors, and RNA‐binding proteins G3BP1 and G3BP2, as well as its proximal localization to the mitochondrial outer membrane where it may facilitate co-translational import into mitochondria111. CLUH also acts as an regulator of PINK1-Parkin mitophagy pathway [G] contributing to mitochondrial network remodelling and health112. Interestingly, mTORC1 is negatively regulated by CLUH within RNA granules, establishing a model where CLUH positively and negatively regulates specific mitochondrial transcripts (such as TFAM and mitoribosomal components)to adapt to nutrient limitation during fasting88. Cluh mRNA was also found to be directly bound by PGC1α during glucagon stimulation in hepatocytes to promote Cluh expression49. This provides a novel layer to PGC1α nuclear function outside classical transcriptional regulation that may connect cytosolic translation of nuclear-encoded mitochondrial transcripts through modulating CLUH mRNA stability or processing during fasting49.

Brown adipose tissue also employs unique mechanisms to regulate mitochondrial gene expression using FAM195A, a recently identified RNA-binding protein that contains an extended disordered low-complexity domain93. These types of proteins are intrinsically disordered and facilitate phase separation within RNA granules94. FAM195A was found to stabilize mRNAs essential for oxidation of branched-chain amino acids [G] (BCAAs) to drive thermogenesis in brown adipose tissue93. Accordingly, knockout mice exhibit whitening of brown adipose tissue, increased sensitivity to cold, and defects in BCAA oxidation signified by increased plasma levels of leucine, valine, and isoleucine93. These initial observations indicate that one mode of translational regulation of mitochondrial mRNAs is through stabilization within RNA granules, emphasizing the need to further study these organelles and their connection to the translational machinery.

Mitochondrial protein translation regulation in respiratory chain assembly

The mitochondrial translation machinery evolved distinctly from its cytosolic counterpart owing to its bacterial origin and subsists to synthesize 13 polypeptides (8 in yeast) that remain encoded in the mtDNA. A synchrony between the mitochondrial and nuclear genomes and surveillance of the mitochondrial proteome via intercompartmental communication and signalling pathways contributes to respiratory chain formation. One approach is through nuclear encoded auxiliary factors that stimulate translation initiation and progression. In yeast, translational activators specific for certain mitochondrial-encoded mRNAs promote translational capacity in mitochondria. For example, there are at least seven translational regulators of complex IV subunits CO1, CO2, and CO395. Translation of mitochondrial mRNAs is predominantly regulated in a positive manner through interaction with the mitochondrial ribosome and localization to specific sites on the inner membrane95. Metabolic cues associated with mitochondrial biogenesis such as a switch in nutrient source away from glucose feedback to translational activators to facilitate the efficient assembly of respiratory complexes95. Furthermore, translational rather than transcriptional programs are synchronized between the cytosol and mitochondria where intercompartmental communication is unidirectional from the cytosol to mitochondria27. This enables yeast mitochondrial translation to swiftly coordinate with levels of synthesis of nuclear-encoded respiratory chain subunits.

Rapid communication between translational machineries does not occur in mammalian mitochondria28. Mitochondrial translational activators are largely missing and mitochondrial RNAs also lack 5’ UTRs present in yeast, requiring a different mode of translational regulation96. TACO1 is the only known regulator of mitochondrial translation and facilitates the translation of mitochondrial CO1 synthesis through binding to various regions of the coding region97,98. Mitochondrial protein homeostasis in higher eukaryotes is also challenged by the fact that translation of subunits within the multimeric respiratory complexes does not occur in stoichiometric balance28,99; there is a 2–32-fold variance in subunit synthesis rates for dual origin (nuclear and mitochondrial genomes) respiratory chain complexes28. Despite this, average protein synthesis for each respiratory complex is highly correlated across compartments to conceivably avoid proteotoxic stress in the mitochondria28. One mediator of this interaction is the mitochondrial RNA binding protein LRPPRC28. In LRPPRC knockout cells, cytosolic and mitochondrial translation are anti-correlated, which induces protein quality control pathways such as the unfolded protein response [G] (UPR) of the ER28. Metabolic stimuli such as growth signalling or nutrient deprivation regulate LRPPRC100,101, suggesting that metabolic signals may establish translational coordination between compartments to drive respiratory chain assembly. For example, mTORC1 controls LRPPRC protein levels in cells101, while the mitochondrial NAD+-dependent deacetylase SIRT3 targets LRPPRC to promote respiratory chain function in fasted livers100.

Translational synchrony occurs on a longer timescale in higher eukaryotes and hinges on the mitochondrial translation machinery sensing the import and assembly of nuclear-encoded respiratory factors and adjusting mitochondrial translation rate accordingly28,95. This mode of feedback regulation allows for fine-tuning of respiratory complex formation and mitigates protein folding stress from unassembled protein subunits. Aside from LRPPRC, assembly factors C12ORF62, MITRAC12, and MITRAC15 adjust synthesis of mitochondrial-encoded CO1 and ND2 with import of nuclear-encoded polypeptides2931,102. TIM21, a component of the mitochondrial protein translocase also mediates the assembly of pre-sequence-containing subunits into early assembly intermediates upon their import into mitochondria31. Although different mechanisms regulate cross-compartment protein synthesis in yeast and humans, it will be important to further define these processes and upstream factors controlling their communication to enhance mitochondrial respiration.

Post-translational mechanisms

Transcriptional and translation control is energetically costly and time-consuming for cells. In order to provide flexibility to environmental cues, cells developed strategies to regulate mitochondrial protein activities which include post-translational modifications (PTMs) and protein-protein interactions. These mechanisms can control mitochondrial protein import, respiratory chain assembly and rapidly adapt mitochondrial function to cellular requirements during acute energetic stress, in contrast to prolonged transcriptional and translational programs. Communication through signalling cascades between the ER and mitochondria including PTM regulation of OMM proteins and Ca2+ influx to regulate TCA enzymes103 also fine-tunes mitochondrial function to cellular demands. In this section, we discuss different mechanisms by which mitochondrial function can be modulated through post-translational mechanisms (Fig. 4).

Figure 4 |. Post-translational mechanisms governing respiratory control and the role of mitochondrial membrane dynamics.

Figure 4 |

Multiple interconnected levels of post-translational control, involving inter-organelle crosstalk, promote mitochondrial respiration. Under certain metabolic conditions such as thermogenesis in beige/brown adipose tissue, exercise or insulin/IGF1 stimulation of muscle tissue, or nucleotide limitation in cancer cells, phospholipid (PL) synthesis of cardiolipin (CL), phosphatidylethanolamine (PE), or ether lipid PE are augmented to facilitate respiratory chain stability, activity, and supercomplex assembly. Synthesis of ether lipids initiates in the peroxisome and terminates in the ER, linking these organelles to mitochondrial respiratory function. Stoichiometric shifts in respiratory chain complexes towards superassembly occur under physiological conditions such as exercise and cellular stress including glucose or nucleotide limitation to enhance mitochondrial respiration. Increased respiratory capacity of mitochondria is also promoted by cristae remodelling, which is supported by the mitochondrial contact site and cristae organizing system (MICOS). MIC60 subcomplex promotes the invagination of the inner mitochondrial membrane (IMM), which is facilitated by stabilization of the membrane via outer mitochondrial membrane (OMM) components such as SAM50. MIC10 subcomplex promotes cristae elongation. The coordination of the MICOS subcomplexes relies on the MICOS component MIC19, which bridges MIC60 and MIC10 subcomplexes. Along with the MICOS complex dynamin-like GTPase and mitochondrial fusion factor, OPA1, maintains cristae junction stability and cristae architecture15,16. Formation of complex V (CV; ATP synthase) dimers are an important structural element at the tip of cristae that provides proper curvature213,216,256. Mitochondrial–ER contacts (MERCs) — where OMM is in close apposition to the ER membrane — regulate various aspects of mitochondrial structure and function, including protein translation and import, lipid transport, membrane dynamics and Ca2+ signalling. During ER stress (which can result from, for example, glucose deprivation or cold stimulation), ER-localized PERK kinase activates O-GlcNAc transferase (OGT), which results in O-GlcNAcylation of the receptor TOM70. This promotes TOM70-dependent protein import, including import of MIC19, supporting cristae formation and oxygen consumption (at least in brown adipocyte mitochondria). Cristae biogenesis is also correlated with respiratory chain superassembly. Mitochondrial fusion and fission also impact mitochondrial respiratory capacity, tying with regulation of the assembly and organization of respiratory complexes and cristae dynamics. Fusion–fission events are themselves regulated by lipid remodelling. PA, phosphatidic acid; PC, phosphatidylcholine; PG, phosphatidylglycerol; PI, phosphatidylinositol; PS, phosphatidylserine; DRP1, dynamin-related protein 1; MFN1, mitofusin 1.

Respiratory chain complex regulation by PTMs

PTMs incorporate chemical groups to amino acid side chains that modulate the activity and assembly of proteins and protein complexes. The respiratory machinery of mitochondria comprises several complexes with multiple subunits that are subject to different modifications104. The role of these modifications and their regulation needs to further study, but their stoichiometries are high enough to be detected by cryogenic electron microscopy105 highlighting their potential importance in regulating OXPHOS activity. In this subsection we summarize the most relevant modifications on OXPHOS complexes and their impact on mitochondrial respiration.

Phosphorylation is most widely studied modification across the proteome. More than 90% of mitochondrial proteins contain at least one known phosphorylation site104. Phosphorylation is dependent on OXPHOS activity which yields the ATP necessary for kinase reactions106. Respiratory complexes are phosphorylated at multiple sites with distinct results on activity and assembly. Protein Kinase A (PKA) has been postulated as a major kinase for different mitochondrial proteins107111. Phosphorylation of respiratory complex I (NDUFA1, NDUFS4 and NDUFB11) by PKA increases the assembly and function of this complex107,108,112. PKA also phosphorylates respiratory complex IV subunit COX4I1 on serine 58 which is particularly relevant for more efficient electron flux under conditions where cells are reliant on OXPHOS activity109,110. Phosphorylated Ser58 in COX4I1 prevents binding and allosteric inhibition by ATP, and toggles between activation (phosphorylated) and inactivation (non-phosphorylated) of the ETC110. Under hypoxic conditions, PKA can inhibit complex IV activity by phosphorylation on Ser40 to prevent excessive ROS production113. Other kinases also regulate the function of different OXPHOS subunits. The cyclin B1–Cdk1 complex phosphorylates several components of complex I including NDUFA12 (on Thr120 and Thr142), NDUFB6 (Thr5, Ser29 and Ser55), NDUFS2 (Ser364), NDUFV1 (Thr383), and NDUFV3 (Ser105) which promotes complex I activity during mitosis114. PINK1 phosphorylates NDUFA10 on Ser250 phosphorylation to augment complex I activity115, and mitochondrial c-Src kinase phosphorylates Tyr215 of succinate dehydrogenase complex II subunit SDHA which is required for electron transfer and prevention of ROS production116. Fgr tyrosine kinase-dependent phosphorylation of Tyr604 increases complex II activity and compensates for complex I inhibition117,118. While the landscape of protein phosphorylation in the OXPHOS complexes is abundant, further research is necessary to determine the responsible kinases and understand their implications in mitochondrial respiratory function.

Acetylation is the addition of acetyl groups to lysine amino groups. This process can be regulated by acetyl-transferases, but due to the high concentrations of acetyl-CoA inside mitochondria (>1 mM) and slightly alkaline conditions (pH 8.0) in the matrix, mitochondrial proteins can spontaneously be acetylated119. The general regulation of protein acetylation has been previously reviewed120 and here we focus on the impact on respiratory function. Deacetylase sirtuin 3 (SIRT3) counteracts the effects of acetylation on multiple mitochondrial complexes including the OXPHOS components121. Depletion of SIRT3 has revealed decreased activity of complex I in vivo and in vitro and associates with poor recovery of hearts after ischemia122125. Similarly, complexes II and V are deacetylated by SIRT3 to increase their activities125,126. Of note, most of these works manipulate SIRT3 which deeply alters mitochondrial function127 and therefore it is necessary to uncouple the OXPHOS specific effects from the general mitochondrial dysfunction.

Cysteine oxidation from ROS is an important regulatory modification on cellular and specifically, mitochondrial proteins128. In brown adipocytes, UCP1 dissipates the proton gradient across the mitochondrial inner membrane. Under conditions of elevated glucose, fatty acid, or branch chain amino acid oxidation, UCP1 Cys253 becomes oxidized, further increasing its activity, mitochondrial respiration, and thermogenic capacity in this tissue10. Abolition of Cys253 oxidation results in obesity and inflammation129. Another example of ROS-dependent respiratory regulation is the oxidation of Cys39 in ND3 of complex I130132. Oxidation of this residue prevents ROS production due to reverse electron transfer (i.e: reductive activity of complex I from ubiquinol to NADH formation) during ischemia.

Mitochondrial protein import and respiratory control

Mitochondrial biogenesis relies on the existence of mitochondrial import machineries that are responsible for the import of most nuclear-encoded mitochondrial proteins. Consequently, precursor protein import is a critical and highly regulated process that ensures the proper assembly of membrane protein complexes required for cristae formation and the respiratory chain biogenesis. The TOM complex at the outer mitochondrial membrane is the main entry gate for protein import133. The main core of the TOM complex (also known as General Import Pore, GIP) consists of subunits TOM40, TOM22, and the small components TOM5, TOM6, and TOM7. TOM40 has a distinct β-barrel structure as opposed to the α-helical structure of the other subunits and interacts with N-terminal positively charged fragments of precursor proteins to facilitate their import into the intermembrane space134136. Aside from the GIP, the TOM complex also includes TOM20 and TOM70 receptors that recognize and anchor precursor proteins prior to their import through the TOM40 pore.

TOM70 and TOM20 receptors are essential to recognize and drive precursors to the GIP. TOM20 participates in the recognition of mitochondrial precursors that contain a pre-sequence motif (ϕχχϕϕ, where ϕ is hydrophobic residue and χ is any amino acid) while TOM70 has been classically associated with the import of mitochondrial ion and metabolite carriers137141. Further studies have expanded the range of TOM70 substrates showing the ability of the receptor to bind non-canonical substrates142 including hydrophobic140,141 and aggregation-prone proteins143, and proteins with internal mitochondrial targeting sequence-like signals [G] (iMTS-L)144. Moreover, TOM70 and TOM20 can participate in the import of identical substrates in a sequential manner where TOM70 transfers to TOM20 before reaching the GIP145, demonstrating a crosstalk between the TOM components to regulate mitochondrial protein import. This may explain the compensatory effects on protein import of TOM70 substrates when TOM70 is absent143,145, preventing the identification of bona fide substrates. Mutations in the receptor TOM70 negatively affect the assembly of respiratory complexes showing direct connection between this receptor and mitochondrial respiration146 although the specific mechanism on complex assembly is unclear. The delivery of precursors from TOM70 to the GIP is still under debate. Direct interactions with TOM70 and mitochondrial precursors have been shown in vitro147 and may be mediated by Cys-Cys disulfide bonds148. However, since TOM70 interacts with heat shock proteins 70 (HSP70) and 90 (HSP90)140,141, direct delivery of mitochondrial protein precursors from these chaperones to the GIP has been proposed similar to the SEC translocon [G] systems in the ER149. Transient interactions between the chaperone clamps of TOM70 and the EEVD domain of HSPs increase the concentrations of chaperone–precursor complexes at the gate of the GIP to drive import into mitochondria140,149. The involvement of TOM70 in immune responses150,151 and Ca2+ signalling152 expand the functions of the receptor beyond protein import and suggests a central role in cellular signalling.

Regulation of TOM complex activity is essential to ensure proper mitochondrial function153. PTMs control TOM complex activity. Initial studies in yeast established that phosphorylation events control both the biogenesis the of TOM complex and precursor protein import through cytosolic kinases Ck2 and Pka137. Ck2 phosphorylates the receptors Tom22 and Tom20, which promotes biogenesis of the TOM complex137. In contrast, Pka-driven phosphorylation of Tom70 impairs the binding and import of, at the very least, classical yeast Tom70 substrates AAC carrier [G] and PiC carrier [G]137. In mammals, an interplay between O-GlcNAcylation and phosphorylation at Ser94 of TOM70 controls mitochondrial protein import that is linked to cristae formation during cold stress in brown adipocytes21. Upon norepinephrine (NE) or cold stress, activation of the ER kinase PERK triggers a post-translational modification cascade that results in increased TOM70-dependent precursor protein import. More specifically, PERK-activated protein O-GlcNAc transferase (OGT) directly and indirectly stimulates precursor protein import through glycosylation of TOM70 and CK2, respectively, driving mitochondrial cristae formation and cellular respiration in brown adipocytes21. CK2-driven phosphorylation negatively affects TOM70 activity and is prevented by competition with O-GlcNAc21,154,155 occurring on the same site and by OGT-mediated inhibition of CK2 activity155. Previous studies found that CK2 negatively controls brown adipocyte function and mitochondrial biogenesis156 and inhibits PERK activities157 critical for ER–mitochondrial communication158, reinforcing the importance of this signalling network on mitochondrial function. Given these results, PTMs of the TOM complex are a conserved method to regulate precursor protein import, however, the specific PTMs and cytosolic enzymes involved adapted to different cellular functions and needs.

Mitochondrial protein import functions in concert with numerous ATP-dependent chaperones and AAA+ proteases [G] that ensure proper import, processing, and stabilization of protein substrates159. The mitochondrial matrix chaperones HSP60 and mitochondrial HSP70 (mtHSP70) fold precursor proteins to avoid protein aggregation159. TIM23-dependent protein import across the mitochondrial inner membrane and subsequent protein folding is mediated by mtHSP70 (ref.159). In conjunction with mtHSP70, the matrix protease LONP acts as a chaperone to promote protein incorporation into the inner membrane and matrix independent of its protease function160162. LONP anti-aggregation activity acts towards numerous substrates including subunits of the respiratory chain, mitochondrial ribosome160,161, and the OXA1Linsertase of the inner membrane161. Mitochondrial proteases are also downstream targets of cellular signalling pathways. In response to hypoxia or serum starvation in cancer cells, LONP is phosphorylated by AKT [G] to enhance its protease activity and limit accumulation of unfolded respiratory chain subunits163. The i-AAA protease YME1L is hyperactivated under hypoxic or glutamine starvation conditions, degrading mitochondrial protein translocases and lipid transfer proteins to limit mitochondrial biogenesis164. Altogether, the tight control of the import machineries and processing of mitochondrial protein precursors by proteases165 is a limiting step in the regulation of all mitochondrial processes and a quality control166 mechanism that responds to cellular stressors such as hypoxia, fasting, or cold.

Assembly of mitochondrial respiratory protein complexes and supercomplexes

During assembly of the respiratory chain, intermediate subcomplexes and a growing number of accessory proteins progressively interact to form intact respiratory complexes over the time course of hours167,168. Once assembled, respiratory complexes dynamically form higher order assemblies termed supercomplexes in response to metabolic pressures. Complex I predominantly exists in complex with other respiratory complexes including a dimer of complex III (I+III2; subscript denotes stoichiometry) or in conjunction with complex III and IV (SC I+III2+IV); complex III exists as an obligate dimer and in complex with complex IV (SC III2+IV1–2); complex IV exists mainly as a monomer, but to a lesser extent as dimers and multimers169171. The co-existence of isolated complexes and their super-assemblies is known as the plasticity model of the electron transport chain172,173, which asserts a functional role of varied assemblies on electron transport. Consistent with this scenario, stoichiometry of isolated and superassembled respiratory chain complexes is regulated in cells under the physiological stimuli of exercise in muscle170,174 and fasting in liver175.

The physiological role of supercomplexes has been heavily debated with the main models contending reduced ROS production176,177, CoQ and CytC substrate channelling [G]175,177, enhanced kinetics178181, respiratory chain complex stability182184 and assembly185, and diminished protein aggregation in the protein-rich inner mitochondrial membrane186. Despite this, metabolic and physiological conditions that modify stoichiometries of supercomplexes to isolated complexes are consistent with increased efficiency or quantity of mitochondrial respiration. Supercomplexes are elevated in proliferative cells in response to changes in nutrient conditions or metabolic constraints18,20,187. Cells facing ER stress from glucose deprivation upregulated the PERK–eIF2α–ATF4 pathway to transcriptionally induce supercomplex assembly factor 1 (SCAF1), which acts a bridge between CIII2 and IV to promote supercomplex III2+IV (a.k.a. Q-respirasome) formation178,188,189 as well as particular I+III2+IV supercomplexes (also known as N-respirasomes)188,189,187. This resulted in enhanced cellular respiration, ATP production, and proliferation20. Increased N-respirasome assembly driven by lipid remodelling in the mitochondria (see also next section) also supports growth and proliferation of cancer cells under nucleotide limitation18. Furthermore, supercomplexes promote tumour growth under oxygen limited conditions190,191 and the proliferation and viability of acute myeloid leukaemia cells in vivo192. These findings point to a model where supercomplexes provide metabolic flexibility under metabolically demanding conditions to fuel proliferation and highlight the therapeutic potential of targeting supercomplexes or their assembly factors in cancer. Proof of this model was acquired in a recent study of the yeast Q-respirasome, where mutations in the complex III-subunit Cor1 that specifically impaired supercomplex assembly, limited the diffusion of cytochrome c between complex III and IV and decreased coupled respiration with NADH and succinate as electron donors180. This led to a competitive disadvantage of mutant strains cultured in nutrient conditions that require mitochondrial respiration, which was rescued through cytochrome c overexpression180. Consistent with this, cryo-EM kinetic studies established that electron transfer between complex III and complex IV in the yeast Q-respirasome involves 2D diffusion of positively charged cytochrome c across the negatively charged surface of the supercomplex, in contrast to bulk exchange with solvent (3D diffusion)179. This charged surface is conserved in the mammalian Q-respirasome and may explain the enhanced respiratory activity of superassembled complex III and IV178.

Supercomplexes control metabolic activity under energetically demanding conditions in organismal physiology as well. In humans and mice, endurance exercise training causes a redistribution of respiratory complexes into supercomplexes that correlates with increased mitochondrial respiration170,174. This appears to be functionally relevant since wild-type SCAF1 male and female mice exhibit higher top speeds in exercise performance compared to their mutant or knockout counterparts177. SCAF1 in zebrafish also promotes growth, increases female fecundity, and prevents fat deposition193, while in mice it facilitates bodyweight maintenance of male mice under severe intermittent fasting177. Thus, respirasomes appear to increase the efficiency of mitochondrial respiration using certain substrates, a subtle phenotype that becomes apparent under specific metabolic or physiological states. However, they can also be selected against to optimize mitochondrial respiration. For example, N-respirasomes are reduced when certain fuel sources with low NADH/FAD electron ratios are preferred such as fatty acids in liver during fasting175.

Mitochondrial membrane remodelling

Mitochondrial membranes represent an intricate network of protein-protein and protein-lipid interactions that dictates mitochondrial morphology and functions such as respiration18,20, TCA cycle activity, and metabolite trafficking17. The composition of mitochondrial membranes is subject to continuous change and altered by protein translation28,95 and lipid synthesis rates18, which depend on metabolic cues. We previously discussed the importance of protein translation and mitochondrial protein translocation in sustaining mitochondrial respiratory function. Here, we discuss the regulatory mechanisms that control mitochondrial membrane morphology and dynamics and consider how these processes impact respiratory function of mitochondria.

Regulation of mitochondrial membrane dynamics

Mitochondria are highly dynamic structures that undergo continuous cycles of membrane fusion and fission to expand or contract the mitochondrial network. These opposing events maintain mitochondrial network homeostasis and adjust mitochondrial function to different cellular conditions such as nutrients, oxidative stress, or other internal/external signals. In general, membrane fission events help recycle damaged mitochondria via mitophagy, while fusion events distribute mitochondrial contents including proteins and mtDNA to prevent accumulation of pathological molecules22. Altogether, these events safeguard the health of the mitochondrial network and ensure optimal respiratory capacity. As such, an imbalance in mitochondrial membrane dynamics leads to cellular dysfunction and pathology, in particular affecting the nervous system, perhaps reflecting the high dependency of neurons on mitochondrial respiration194. Membrane fission and fusion events are controlled by dynamin-like GTPase proteins and their accessory factors, which are excellently reviewed elsewhere22,195. Mitochondrial fission requires PTM-mediated dynamin-related protein 1 (DRP1) localization to mitochondria and occurs at sites where ER tubules constrict mitochondria196. During glucose or glutamine starvation or ischemic conditions, PKA-mediated phosphorylation of DRP1 prevents its localization to mitochondria and leads to unopposed mitochondrial fusion, cristae abundance, maintenance of ATP levels, and cell viability197199. Cell proliferation and oncogenic transformation also stimulates mitochondrial fusion, respiration, and ATP output200. Mitochondrial fragmentation is also associated with increased respiration in certain physiological contexts such as thermogenic/adrenergic stimulation of brown adipocytes201, exercise in cardiac tissue202, and immune activation of B and T cells203,204. Thus, general rules about the relation between mitochondria fission/fusion dynamics and enhanced respiratory rates may not exist. Instead, mitochondria uniquely adapt their morphology to different cellular stimuli in order to promote mitochondrial respiration.

Independent of general length and connectivity changes to the mitochondrial network, the mitochondrial inner membrane undergoes structural rearrangements forming cristae. Cristae are invaginations of the inner mitochondrial membrane that harbour the respiratory complexes1 and sustain respiratory super complex assembly and function205. Cristae are a structural platform that allows proper distribution and function of respiratory complexes to maximize respiratory capacity20,205. For instance, cold stimulation in brown adipocytes stimulates cristae formation to increase respiratory function without affecting levels of OXPHOS components21,206. MICOS, OPA1 and complex V dimers along with specific lipid composition shape cristae architecture. The MICOS complex consists of two distinct subcomplexes, MIC60 and MIC10, that provide spatial organization for cristae14,207. OPA1 also plays a role in the stabilization of cristae junctions, tying in cristae dynamics with the formation of respiratory supercomplexes1416. Under low nutrient conditions, associations between OPA1 and mitochondrial solute carriers modulate OPA1 function which can regulate complex V assembly and cristae function and morphology, independently of its fusion-related GTPase activity208. Cristae formation is initiated at the cristae junction in a process where the MIC60 subcomplex promotes the invagination of the inner mitochondrial membrane14. This process is also favoured by the stabilization of the mitochondrial membranes by interactions between the MICOS and outer mitochondrial membrane components such as SAM50 (ref.209). Later, the MIC10 subcomplex promotes cristae elongation14. Despite the predominant role of each MICOS subcomplex in different aspects of cristae biogenesis, both interact and are not independent entities. For example, loss of MIC60, MIC19, MIC13 or MIC10 is sufficient to cause dramatic defects in cristae density and morphology14,210,211. The coordination of the MICOS subcomplexes relies on the MICOS component MIC19, which bridges MIC60 and MIC10 subcomplexes in yeast212. Similarly, mammalian MIC19 is a limiting component4 that supports cristae biogenesis through the stabilization of inner and outer mitochondrial membrane interactions209. Respiratory complex V abundance is correlated with cristae formation213215. Complex V dimers increase the flexibility of cristae curvature216 at the edge of cristae along with mitochondrial lipids such as cardiolipin217. However, lower levels of complex V do not necessarily imply reduced cristae density. An example of this is brown adipose tissue, where complex V is clearly underrepresented (owing to prevalence of uncoupled respiration in this tissue)218 and dense and highly functional cristae are still formed21.

The role of the MICOS subcomplexes and OPA1 in cristae formation are extensively reviewed219, however, regulation upstream of MICOS components is not well defined. The key player is the ER, which communicates with the mitochondria directly via membrane contact sites, known as MERCs to regulate various aspects of mitochondrial dynamics (Box 2). In proliferative cells, ER stress kinase PERK controls cristae formation during nutrient stress conditions through an unknown mechanism20. In post-mitotic brown adipocytes, the PERK–OGT–TOM70 axis (see subsection Mitochondrial protein import and respiratory control) controls mitochondrial import of MIC19 and cristae formation21. Consistent with a coordination between mitochondrial protein import and cristae biogenesis, TOM70 co-localizes with MICOS complexes at the cristae junction19. MICOS subunits contain a Cys-X9-Cys twin motif (CHCH domain) that increases protein instability and aggregation propensity210,220. This cysteine motif may justify the requirement of the TOM70 pathway as a chaperone to assist the recognition148 and import of MIC19 (refs.21,221). Interestingly, loss of MIC19 can be compensated by overexpression MIC25 (ref.209), a highly similar MICOS component222 with lower endogenous expression levels. Compared to other MICOS components, MIC19 import seems insensitive to membrane potential fluctuations21,221,223 and therefore explains its pivotal role in cristae organization and stability especially under stress conditions such as norepinephrine stimulation where membrane potential decreases with increased respiration.

Box 2 |. Mitochondrial-ER contacts and mitochondrial functions.

A crosstalk between the endoplasmic reticulum (ER) and mitochondria regulates mitochondrial dynamics. Communication between these two organelles controls fission and fusion events22, Ca2+ release12 and cristae formation21. Lack of mitochondrial–ER communication is associated with the development of age-related diseases including obesity265 and type 2 diabetes266.

Physical contact between the ER and mitochondria constitutes the Mitochondrial-ER Contacts (MERCs). MERCs provide a tethering system that stabilizes communication between organelles. Mitofusins are proteins anchored at the outer mitochondrial membrane and the ER that actively participate in the formation of MERCs267. Mitofusins interact with each other267 and have been shown to establish contacts with ER stress sensors such as PERK268, which is an important regulator of ER stress responses, mitochondrial protein import and cristae formation21. Mitochondrial fission and fusion events occur at the MERCs where mitofusin 1 (MFN1), MFN2 and Mitochondrial Dynamin Like GTPase OPA1 actively participate to define mitochondrial dynamics196,269.

MERCs regulate Ca2+ trafficking between the ER and mitochondria. Ca2+ entry into mitochondria defines cellular fitness and the ability to generate ATP by the OXPHOS complexes. Ca2+ activates pyruvate dehydrogenase, oxoglutarate dehydrogenase, and isocitrate dehydrogenase, which increase NADH levels and ETC activity12,270. Ca2+ is provided by the ER to mitochondria through the ER component inositol 3-phosphate receptor (IP3R) which forms a complex with DJ-1, GRP75, and mitochondrial voltage-dependent anion channel (VDAC) to regulate the transfer of Ca2+ from the ER to the mitochondrial matrix via the mitochondrial calcium uniporter (MCU)271. ER-resident vesicle-associated membrane protein B (VAPB) interacts with outer mitochondrial membrane protein tyrosine phosphatase-interacting protein-51 (PTPIP51) to promote the assembly of a tether complex that allows MERC formation272. Loss of this interaction results in improper Ca2+ trafficking and mitochondrial dysfunction272274. Other components such as mitochondrial import receptor TOM70 of the outer membrane152 and mitofusins275 are implicated in ER-mitochondrial Ca2+ trafficking via interaction with IP3R and stabilization of MERCs, respectively.

Lipid trafficking between the ER and mitochondria is an open paradigm. The ER is the major supplier of lipids for mitochondria by means that are nebulous. Aside from vesicular lipid diffusion between organelles, the ERMES system in yeast225 and VPS13D in mammals226 participate in the transfer of lipids to mitochondria, indicating an enzymatic-driven machinery that can dictate lipid fluxes to adapt mitochondrial structure (e.g.: cristae biogenesis) and respiratory function. MERCs are a central node for lipid trafficking between the ER and mitochondria.

MICOS components are also controlled by PTMs. PKA-dependent phosphorylation of MIC60 downregulates PINK1 and subsequent recruitment of Parkin to the mitochondria, negatively regulating mitophagy/mitochondrial clearance111. In Drosophila, PINK1 mediates the phosphorylation of MIC60 and sustains cristae formation through MIC60 oligomerization which positively controls cristae formation and neuronal function224. Together, these connections of MIC60 with PINK1 and Parkin111,224 hint at a connection between cristae formation and mitochondrial function that prevents oxidative stress via efficient respiration and mitophagy. PINK1 phosphorylation sites in Drosophila seem to be conserved across species224, while PKA-dependent sites appear to be exclusive of mammals111. The presence of different phosphorylation sites might be a consequence of divergent evolutionary adaptive responses, however, the interplay between PKA and PINK1 indicates a quality control mechanism that dictates inner mitochondrial membrane dynamics through post-translational control of the MICOS components.

Cristae biogenesis requires a balance between protein–protein interactions and lipid incorporation into cristae membranes. The balance and contribution of these processes is an open question. Mitochondrial–ER contacts establish a node of communication, which would allow for lipid traffic to feed mitochondrial membrane formation, yet the contribution of this lipid transfer at membrane contact sites as well as proteins involved in this process are largely unknown. In yeast, the ERMES system consisting of Mmm1, Mdm10, Mdm12, and Mdm34 is the primary candidate for tethering non-vesicular transfer of lipids between ER and mitochondria225. In mammals, VPS13D has been identified as a lipid transfer protein at ER–mitochondria contacts226. Unlike molecule diffusion, the involvement of enzymes in lipid trafficking denotes specificity and efficiency for these processes. On the mitochondrial side, the MICOS complex is implicated in mitochondrial lipid uptake. MIC26 and MIC27 which are members of the MIC10 subcomplex participate in the trafficking of incoming lipids from the ER across the mitochondrial membranes, which could support cristae formation and respiratory function227,228. Loss of MIC26 or MIC27 has little effect on cristae biogenesis as opposed to point deletions of MIC60, MIC19, MIC13, or MIC10 downregulation14,210. However double deletion of MIC26 and MIC27 negatively impacts cristae formation and respiratory complex formation perhaps due to a compensatory expression of one of the subunits when the other is downregulated227. The study of mitochondrial–ER protein interactions as well as MICOS assemblies and lipid uptake by mitochondria is fundamental to understand cristae formation and dynamics, and in extension, to understand the regulation of mitochondrial respiratory function.

Mitochondrial lipid composition and remodelling

Membranes are composed of different phospholipid species together with other membrane lipids such as sphingolipids [G] and sterols. The physiochemical properties of membranes including viscosity, permeability, curvature, thickness, tension, and lateral compartmentalization control protein dynamics and signalling cascades. Phospholipids contain a phosphate headgroup modified by choline, ethanolamine, or other organic molecules and fatty acid chains with varying degrees of length and saturation that are connected through a glycerol backbone. Cylindrical phospholipids like phosphatidylcholine form bilayers and create liquid crystalline lamellar phases, while others that possess a small headgroup in comparison to their acyl chains like phosphatidylethanolamine are conical and produce non-bilayer structures such as inverted hexagonal phases229231. Lipid composition differs considerably across organelles, but is fairly consistent for a particular organelle across cell types, highlighting the importance of phospholipid ratios231. Mitochondrial membrane composition is tailored to each mitochondrial compartment to enable membrane shape and the function of protein machineries within these locations. The conical lipids cardiolipin and phosphatidylethanolamine in the mitochondrial inner membrane control the assembly, stability, and activity of respiratory chain complexes, and promote the higher order assembly of mitochondrial supercomplexes that alter the efficiency of the electron transport chain232235. Phospholipid synthesis, transport, remodelling, and degradation are all means to alter mitochondrial phospholipid composition in response to physiological stimuli such as cold exposure or other metabolic stressors17,18,236239.

The connection between membrane composition and metabolic stimuli or growth signalling pathways in eukaryotes is accepted, but not highly explored. Initial observations of yeast cultured on a non-fermentable carbon source such as lactate found that levels of mitochondrial phosphatidylethanolamine and cardiolipin increased to aid respiratory chain assembly240,241. In mammals, thermogenic capacity is modulated through activation of beige and brown adipose depots and requires changes to mitochondrial phospholipid composition to support increased electron transport activity and consumption of metabolic substrates including glucose, lipids, and BCAAs17. Mice housed at 5°C displayed specific induction of proteins involved in phospholipid metabolism that caused dramatic accumulation of both phosphatidylglycerol, a precursor in cardiolipin synthesis, and nascent cardiolipin species17. Accordingly, adipose-specific cardiolipin synthase 1 (Crls1) knockout mice exhibited decreased mitochondrial mass, abnormal cristae structure, reduced respiratory chain complex and supercomplex assembly and diminished uncoupled respiration/thermogenesis that led to whole-body metabolic defects and insulin resistance17. Humans also accumulate phosphatidylglycerol species such as Lyso-PG 18:0; Lyso-PG 18:1, and PG 20:0/22:5 in serum following 1 hour cold challenge consistent with observations in mice236, underscoring the requirement of brown adipose tissue cardiolipin synthesis to thermogenesis across mammals.

Cancer cells dynamically regulate their lipidome, including that of their mitochondria, to adapt to nutrient limitations and the tumour environment242. One recent demonstration of this is remodelling of mitochondrial lipid composition under nucleotide limitation, which sustains respiration and division in proliferative cells18. Specifically, reduction of pyrimidine synthesis led to the change in the make-up of phosphatidylethanolamine and phosphatidylcholine in the mitochondria, with the substitution of normally occurring diacyl species with their ether-lipid counterparts18. Ether lipids are peroxisome-derived phospholipids where the acyl chain at the sn-1 position [G] is attached through an ether linkage in contrast to an ester bond of diacyl phospholipids243. This structural change leads to the formation of non-lamellar inverted hexagonal structures, which facilitate membrane fusion events and the formation of lipid microdomains243,244. In mitochondria, ether phospholipid accumulation promoted the formation of active respiratory supercomplexes that enabled proliferation under low nucleotide conditions18. This link between mitochondrial lipid composition and respiratory efficiency has also been reported in the context of the Barth syndrome [G], where cells are defective in cardiolipin remodelling and contain low levels of ether phospholipids and reduction in supercomplexes245,246. Regarding carcinogenesis, ether phospholipids promote hypoxia tolerance239 and aggressiveness of certain cancers247249. Thus, modulation of ether phospholipid ratios within cellular membranes to, in part, regulate mitochondrial function, represents a strategy exploited by cancers to adapt to environmental and metabolic stress. A better understanding of their upstream regulatory control will enable us more efficiently target phospholipids in health and disease.

Conclusions and perspective

Mammalian cells use a variety of regulatory mechanisms for mitochondrial respiratory adaptations in response to intracellular or extracellular signals. One of the regulatory levels encompasses transcriptional control of mitochondrial gene expression, which, in large part, depends on chromatin complexes containing PGC1α — a program that promotes mitochondrial biogenesis and increased activity of the respiratory chain, and occurs in physiological adaptations to lower temperatures, physical activity or in certain types of tumours. Another level of control is executed at the translation of mitochondrial transcripts encoded in the nuclear genome, whereby rates of protein synthesis increase steady-state levels of subsets of mitochondrial proteins. Components of this control include the mTORC1–4E-BP1/2 axis as well as the RNA-binding protein CLUH that mediates formation of RNA granules for stability and translation of specific mRNAs. Post-translational regulatory mechanisms include various signalling-dependent protein chemical modifications that modulate protein interactions and activities. Assembly of respiratory complexes and supercomplexes, mitochondrial protein import, MICOS and OPA1-dependent cristae formation, and mitochondrial lipid composition coordinate with PTMs to adjust mitochondrial respiration and satisfy cellular energetic demands. These different levels of regulation are ultimately linked, to allow the cells to respond adequately to various cues. For example, adaptation to lower temperatures in mammalian brown adipocytes synchronizes transcriptional and translational programs of nuclear and mitochondrial encoded genes that reorganizes mitochondrial architecture and metabolic function to ensure adaptive thermogenesis.

Although the general regulatory mechanisms for mitochondrial adaptation are now known in some detail, the future challenge will be to address how these mechanisms operate in different pathological and physiological contexts. Mitochondrial energetic adaptations, such as physical activity or cold exposure, confer significant health benefits including cellular fitness and protection against metabolic diseases (Box 3). The inability of mitochondria to adapt results in chronic bioenergetic defects and cumulative cellular damage that causes inflammation which can contribute to neurodegenerative pathologies (Box 3). To comprehensively understand mitochondrial adaptive mechanisms, essential questions still need to be addressed. For example, the transcription coactivator PGC1α increases mitochondrial respiration and biogenesis. However, it is largely unknown how the induction of mitochondrial proteins by PGC1α in different stoichiometries across multiple tissues drives remodelling of the mitochondrial proteome and ultimately adapts respiration. Another important area of future research includes dissecting the regulatory mechanisms and components that change mitochondrial organelle architecture to stimulate respiration. For instance, a complete understanding on how cristae are formed in different cell and tissue types is unknown. In response to lower temperatures, brown adipocytes increase MICOS complex assembly, but how new cristae membranes arise from coordination of MICOS and phospholipid trafficking from the ER and mitochondria is unclear. At the molecular level, it is still open how mitochondrial protein import machineries, respiratory complexes and mitochondrial cristae components assemble to regulate mitochondrial respiration during adaptive responses in different physiological or pathological contexts. Addressing these questions will provide new mechanistic insights into how exercise, nutrition and environmental stressors adapt cellular respiration to bioenergetic needs with therapeutical implications in metabolic and chronic diseases. Novel approaches combining biochemistry, physiology, cell biology, and emerging high resolution microscopy techniques will be needed to address these questions. Finally, integration of physiological and biochemical data with quantitative proteomics, PTM mapping, and metabolomics will further clarify the mechanisms governing mitochondrial function and adaptive capacity.

Box 3 |. Examples of tissue-specific physiological/pathological implications of defective adaptation of mitochondrial respiratory function.

Mitochondrial respiratory chain activity is required for normal function in tissues with high energy demand such as brain, heart, and skeletal muscle. Diseases caused by underlying mitochondrial dysfunction often affect these organ systems. For example, mitochondrial diseases are genetically inherited disorders caused by mutations in nuclear- or mitochondrial-encoded mitochondrial genes. These mutations impact respiratory chain function either directly (mutation in genes encoding respiratory chain subunits) or indirectly (mutation in genes encoding mitochondrial DNA replication, transcription, translation, or membrane fission/fusion (OPA1/DRP1) machineries). One hallmark of mitochondrial DNA (mtDNA) mutations is exercise intolerance and muscle fatigue, which is characterized by an inability to maintain force during repeated or sustained muscle contractions, resulting from progressive dysfunction of the respiratory chain276. This stems from a mosaic impairment of mitochondrial quality control via autophagy (mitophagy) in muscle fibres, driving accumulation of mitochondria with pathogenic (mutated) mtDNA and concomitant respiratory chain deficiency, resulting from the inability of such mitochondria to generate functional respiratory complexes277.

Impairment of transcriptional regulation of mitochondrial function has considerable pathological implications. In brown fat, the inability to increase mitochondrial genes during cold or overnutrition results in defects on thermoregulation, energy balance, and diabetes278,279. In skeletal muscle, loss of PGC1α affects exercise capacity and/or causes increased fatigue due to the loss of slow twitch muscle fibres, which are highly enriched in mitochondria26,35. Other physiological conditions where mitochondrial energetic adaptation is required are perinatal cardiac adaptation280 or kidney function281. For example, defects in the transcriptional regulation of mitochondrial biogenesis cause acute renal injury and fibrosis282. Certain parts of the brain such as the substantia nigra [G] are especially sensitive to mitochondrial deficits283. Loss of dopaminergic neurons occur in Parkinson disease and is associated with defects in the Parkin-mediated mitophagy regulatory pathway that promotes PGC1α activity284. Other neurodegenerative diseases associated with mitochondrial failures are Alzheimer disease and Huntington disease285. Although less explored, emerging data also implicate the important role of mitochondrial respiratory adaptation in immune cells, such as during T cell or NK cells activation in response to infection or cancer cells286,287.

Defects in other processes indirectly linked with mitochondrial respiratory function also lead to human disease. The mitochondrial phospholipid cardiolipin is essential for the assembly and stability of protein import complexes, metabolite carriers, and the respiratory chain. Consequently, mutations in the cardiolipin fatty-acid-chain remodelling protein TAZ-1 leads to dilated cardiomyopathy, skeletal myopathy, and neutropenia in Barth syndrome, while mutations in cardiolipin biosynthetic enzyme AGK cause hypertrophic cardiomyopathy in Sengers syndrome [G]288. In addition, mutations in mitochondrial protein import machinery such as DDP1/Tim8 and DNAJC19/Tim14 cause the neurodegenerative disorder human deafness dystonia syndrome and dilated cardiomyopathy with ataxia, respectively289. Interestingly, DNAJC19 also participates in cardiolipin remodelling by interacting with prohibitins, a group of mitochondrial proteins that act as scaffolds that define membrane domains288.

Acknowledgements

This work is partially supported by grants from the National Institute of Health (NIH): R01 DK089883 NIDDK, R01 DK081418 NIDDK, R01 DK117655 NIDDK, R01 CA181217 NCI, 9R56 AG074527 NIA, R01 GM121452 NIGMS, and Claudia Adams Barr Award to P.P., F32 GM125243 NIGMS and Charles A. King Trust Postdoctoral Fellowship Program to C.F.B., and Human Frontier Science Program (LT-000033/2019-L) to P.L.M.

Glossary

cristae

Invaginations of the inner mitochondrial membrane that increase the surface area of respiratory reactions and harbour the respiratory complexes

brown and beige adipose tissues

Thermogenic fat cells with abundant mitochondria that oxidize glucose, fatty acids and branched-chain amino acids to generate heat

uncoupling protein 1 (UCP1)

Inner mitochondrial membrane protein expressed in thermogenic tissues that dissipates membrane potential from the respiratory complexes and sustains thermogenesis

Haem

porphyrin group coordinating an iron atom required for electron transfer or oxygen transport

iron-sulfur clusters

iron sulphide molecules that transfer electrons across respiratory complexes

ER stress

a condition that occurs when the capacity of the ER lumen to fold proteins is saturated. ER stress transduces signals to other organelles such as mitochondria to adapt cellular metabolism to satisfy energy demands

mediator complex

a multiprotein complex that transduces signals from transcription factors to RNA polymerase II in order to control gene expression

catecholamines

Group of chemical neurotransmitters including dopamine, epinephrine, and norepinephrine that are released into the blood upon stress and modulate beige and brown fat tissue activity

mTOR

mammalian Target Of Rapamycin, a nutrient-sensor kinase involved in the control of cellular growth, survival, metabolism, and immunity

AMPK

AMP-activated protein kinase that senses fluctuations in ATP/AMP ratio

Translation Initiator of Short 5′ UTR (TISU) elements

Sequence elements that are downstream of transcription start sites and regulate both transcriptional and translational initiation and present in mRNAs with very short 5’ UTRs

PINK1-Parkin mitophagy pathway

a quality control pathway that marks damaged mitochondria to promote their autophagy-mediated destruction

branched-chain amino acids (BCAAs)

amino acids including valine, leucine and isoleucine that can be oxidized in the cell to obtain energy

unfolded protein response (UPR)

A process that regulates a transcriptional and translational response to ER protein folding stress

internal mitochondrial targeting sequence-like signals (i-MTSLs)

peptide sequences present within proteins destined for mitochondria that interact with import receptors and increase import competence

SEC translocon

Protein complex embedded in the endoplasmic reticulum membrane that transports proteins from and through the ER lumen

AAC carrier

ADP/ATP carrier is an inner mitochondrial membrane transporter that exchanges ATP/ADP

PIC carrier

Phosphate carrier, is an inner mitochondrial membrane transporter of phosphate

AAA+ proteases

subset of ATPase proteases that participate in diverse quality control mechanisms in mitochondria and cytosol (26S proteasome)

AKT

(a.k.a., Protein Kinase B, PKB) Group of serine/threonine kinases that respond to a myriad of external stimuli and include AKT1, AKT2 and AKT3

substrate channelling

Biochemical phenomena whereby the intermediate product from one enzyme is shuttled as a substrate for the next enzyme; e.g.: reduced CoQ from respiratory complex I is transferred and oxidized by respiratory complex III2 to reduce cytochrome C. Reduced cytochrome C is oxidized by respiratory complex IV. As a result of this substrate channelling, protons are pumped to intermembrane space through each complex to perform cellular respiration

sphingolipids

Class of phospholipids containing a sphingosine backbone. Sphingolipids facilitate mitochondrial function by stabilizing respiratory complexes but their accumulation correlates with mitochondrial dysfunction and chronic metabolic diseases such as type 2 diabetes

sn-1 position

First stereochemical position on a glycerol moiety to which a fatty acid is attached

Barth syndrome

A rare X-linked genetic disorder of cardiolipin metabolism that presents with cardiomyopathy, neutropenia and muscle weakness

nuclear receptors

ligand-regulated transcriptional factors that are activated by steroid hormones and other lipid-related molecules

substantia nigra

Basal ganglia structure in the brain that plays important roles in behaviour-reward neuronal programs

Sengers syndrome

rare autosomal condition that courses with myocardiopathy, lactic acidosis, muscle weakness and short life expectancy

Footnotes

Competing interests

The authors declare no competing interests.

References

  • 1.Gilkerson RW, Selker JML & Capaldi RA The cristal membrane of mitochondria is the principal site of oxidative phosphorylation. FEBS Lett. 546, 355–358 (2003). [DOI] [PubMed] [Google Scholar]
  • 2.Deshpande OA & Mohiuddin SS Biochemistry, Oxidative Phophorylation. StatPearls (2020). [PubMed] [Google Scholar]
  • 3.Walker JE The ATP synthase: The understood, the uncertain and the unknown. Biochem. Soc. Trans 41, 1–16 (2013). [DOI] [PubMed] [Google Scholar]
  • 4.Enerbäck S et al. Mice lacking mitochondrial uncoupling protein are cold-sensitive but not obese. Nature 387, 90–94 (1997). [DOI] [PubMed] [Google Scholar]
  • 5.Kazak L et al. A Creatine-Driven Substrate Cycle Enhances Energy Expenditure and Thermogenesis in Beige Fat. Cell 163, 643–655 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Nedergaard J & Cannon B Brown adipose tissue as a heat-producing thermoeffector. Handb. Clin. Neurol 156, 137–152 (2018). [DOI] [PubMed] [Google Scholar]
  • 7.Lowell BB & Spiegelman BM Towards a molecular understanding of adaptive thermogenesis. Nature 404, 652–660 (2000). [DOI] [PubMed] [Google Scholar]
  • 8.Chouchani ET et al. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature 515, 431–435 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Granger DN & Kvietys PR Reperfusion injury and reactive oxygen species: The evolution of a concept. Redox Biol. 6, 524–551 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Chouchani ET et al. Mitochondrial ROS regulate thermogenic energy expenditure and sulfenylation of UCP1. Nature 532, 112–116 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Bock FJ & Tait SWG Mitochondria as multifaceted regulators of cell death. Nat. Rev. Mol. Cell Biol 21, 85–100 (2020). [DOI] [PubMed] [Google Scholar]
  • 12.Giorgi C, Marchi S & Pinton P The machineries, regulation and cellular functions of mitochondrial calcium. Nat. Rev. Mol. Cell Biol 19, 713–730 (2018). [DOI] [PubMed] [Google Scholar]
  • 13.Spinelli JB & Haigis MC The multifaceted contributions of mitochondria to cellular metabolism. Nat. Cell Biol 20, 745–754 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Stephan T et al. MICOS assembly controls mitochondrial inner membrane remodeling and crista junction redistribution to mediate cristae formation. EMBO J. 39, e104105 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Frezza C et al. OPA1 Controls Apoptotic Cristae Remodeling Independently from Mitochondrial Fusion. Cell 126, 177–189 (2006). [DOI] [PubMed] [Google Scholar]
  • 16.Civiletto G et al. Opa1 overexpression ameliorates the phenotype of two mitochondrial disease mouse models. Cell Metab. 21, 845–854 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Sustarsic EG et al. Cardiolipin Synthesis in Brown and Beige Fat Mitochondria Is Essential for Systemic Energy Homeostasis. Cell Metab. 28, 159–174.e11 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Bennett CF et al. Peroxisomal-derived ether phospholipids link nucleotides to respirasome assembly. Nat. Chem. Biol 17, 703–710 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kondadi AK et al. Cristae undergo continuous cycles of membrane remodelling in a MICOS ‐dependent manner. EMBO Rep. 21, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Balsa E et al. ER and Nutrient Stress Promote Assembly of Respiratory Chain Supercomplexes through the PERK-eIF2α Axis. Mol. Cell 74, 877–890.e6 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Latorre-Muro P et al. A cold-stress-inducible PERK/OGT axis controls TOM70-assisted mitochondrial protein import and cristae formation. Cell Metab. 33, 598–614.e7 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Giacomello M, Pyakurel A, Glytsou C & Scorrano L The cell biology of mitochondrial membrane dynamics. Nat. Rev. Mol. Cell Biol 21, 204–224 (2020). [DOI] [PubMed] [Google Scholar]
  • 23.Puigserver P et al. A cold-inducible coactivator of nuclear receptors linked to adaptive thermogenesis. Cell 92, 829–839 (1998). [DOI] [PubMed] [Google Scholar]
  • 24.Wu Z et al. Mechanisms controlling mitochondrial biogenesis and respiration through the thermogenic coactivator PGC-1. Cell 98, 115–124 (1999). [DOI] [PubMed] [Google Scholar]
  • 25.Holloszy JO Biochemical adaptations in muscle. Effects of exercise on mitochondrial oxygen uptake and respiratory enzyme activity in skeletal muscle. J. Biol. Chem 242, 2278–2282 (1967). [PubMed] [Google Scholar]
  • 26.Handschin C et al. Skeletal muscle fiber-type switching, exercise intolerance, and myopathy in PGC-1α muscle-specific knock-out animals. J. Biol. Chem 282, 30014–30021 (2007). [DOI] [PubMed] [Google Scholar]
  • 27.Couvillion MT, Soto IC, Shipkovenska G & Churchman LS Synchronized mitochondrial and cytosolic translation programs. Nature 533, 499–503 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Soto I et al. Balanced mitochondrial and cytosolic translatomes underlie the biogenesis of human respiratory complexes. bioRxiv 10.1101/2021.05.31.446345 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Wang C et al. MITRAC15/COA1 promotes mitochondrial translation in a ND2 ribosome–nascent chain complex. EMBO Rep. 21, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Richter-Dennerlein R et al. Mitochondrial Protein Synthesis Adapts to Influx of Nuclear-Encoded Protein. Cell 167, 471–483.e10 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Mick DU et al. MITRAC links mitochondrial protein translocation to respiratory-chain assembly and translational regulation. Cell 151, 1528–1541 (2012). [DOI] [PubMed] [Google Scholar]
  • 32.Rath S et al. MitoCarta3.0: An updated mitochondrial proteome now with sub-organelle localization and pathway annotations. Nucleic Acids Res. 49, D1541–D1547 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Rensvold JW et al. Complementary RNA and Protein Profiling Identifies Iron as a Key Regulator of Mitochondrial Biogenesis. Cell Rep. 3, 237–245 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Holloszy JO Regulation by exercise of skeletal muscle content of mitochondria and GLUT4. J. Physiol. Pharmacol 59, 5–18 (2008). [PubMed] [Google Scholar]
  • 35.Lin J et al. Transcriptional co-activator PGC-1α drives the formation of slow-twitch muscle fibres. Nature 418, 797–801 (2002). [DOI] [PubMed] [Google Scholar]
  • 36.Lai L et al. Transcriptional coactivators PGC-lα and PGC-lβ control overlapping programs required for perinatal maturation of the heart. Genes Dev. 22, 1948–1961 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Ciron C et al. PGC-1α activity in nigral dopamine neurons determines vulnerability to α-synuclein. Acta Neuropathol. Commun 3, 16 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Jiang H et al. Adult conditional knockout of PGC-1α leads to loss of dopamine neurons. eNeuro 3, ENEURO.0183–16.2016 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Tran MT et al. PGC1α drives NAD biosynthesis linking oxidative metabolism to renal protection. Nature 531, 528–532 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Mutlu B & Puigserver P GCN5 acetyltransferase in cellular energetic and metabolic processes. Biochim. Biophys. Acta - Gene Regul. Mech 1864, 194626 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Luo C, Widlund HR & Puigserver P PGC-1 Coactivators: Shepherding the Mitochondrial Biogenesis of Tumors. Trends in Cancer 2, 619–631 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Dominy JE & Puigserver P Mitochondrial biogenesis through activation of nuclear signaling proteins. Cold Spring Harb. Perspect. Biol 5, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Virbasius JV & Scarpulla RC Activation of the human mitochondrial transcription factor A gene by nuclear respiratory factors: A potential regulatory link between nuclear and mitochondrial gene expression in organelle biogenesis. Proc. Natl. Acad. Sci. U. S. A 91, 1309–1313 (1994). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Brown EL et al. PGC-1α and PGC-1β increase protein synthesis via ERRα in C2C12 myotubes. Front. Physiol 9, 1336 (1–17) (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Handschin C et al. Nutritional regulation of hepatic heme biosynthesis and porphyria through PGC-1α. Cell 122, 505–515 (2005). [DOI] [PubMed] [Google Scholar]
  • 46.Yoon JC et al. Control of hepatic gluconeogenesis through the transcriptional coaotivator PGC-1. Nature 413, 131–138 (2001). [DOI] [PubMed] [Google Scholar]
  • 47.Blättler SM et al. Defective Mitochondrial Morphology and Bioenergetic Function in Mice Lacking the Transcription Factor Yin Yang 1 in Skeletal Muscle. Mol. Cell. Biol 32, 3333–3346 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Scarpulla RC, Vega RB & Kelly DP Transcriptional integration of mitochondrial biogenesis. Trends in Endocrinology and Metabolism vol. 23 459–466 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Tavares CDJ et al. Transcriptome-wide analysis of PGC-1α-binding RNAs identifies genes linked to glucagon metabolic action. Proc. Natl. Acad. Sci. U. S. A 117, 22204–22213 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Schreiber SN et al. The estrogen-related receptor α (ERRα) functions in PPARγ coactivator 1α (PGC-1α)-induced mitochondrial biogenesis. Proc. Natl. Acad. Sci. U. S. A 101, 6472–6477 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Cunningham JT et al. mTOR controls mitochondrial oxidative function through a YY1-PGC-1α transcriptional complex. Nature 450, 736–740 (2007). [DOI] [PubMed] [Google Scholar]
  • 52.Puigserver P et al. Activation of PPARγ coactivator-1 through transcription factor docking. Science 286, 1368–1371 (1999). [DOI] [PubMed] [Google Scholar]
  • 53.Wallberg AE, Yamamura S, Malik S, Spiegelman BM & Roeder RG Coordination of p300-mediated chromatin remodeling and TRAP/mediator function through coactivator PGC-1α. Mol. Cell 12, 1137–1149 (2003). [DOI] [PubMed] [Google Scholar]
  • 54.Brandt N, Dethlefsen MM, Bangsbo J & Pilegaard H PGC-1α and exercise intensity dependent adaptations in mouse skeletal muscle. PLoS One 12, e0185993 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Aguilo F et al. Deposition of 5-Methylcytosine on Enhancer RNAs Enables the Coactivator Function of PGC-1α. Cell Rep. 14, 479–492 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Desautels M & Himms-Hagen J Parallel regression of cold-induced changes in ultrastructure, composition, and properties of brown adipose tissue mitochondria during recovery of rats from acclimation to cold. Can. J. Biochem 58, 1057–1068 (1980). [DOI] [PubMed] [Google Scholar]
  • 57.Chouchani ET, Kazak L & Spiegelman BM New Advances in Adaptive Thermogenesis: UCP1 and Beyond. Cell Metab. 29, 27–37 (2019). [DOI] [PubMed] [Google Scholar]
  • 58.Perry CGR & Hawley JA Molecular basis of exercise-induced skeletal muscle mitochondrial biogenesis: Historical advances, current knowledge, and future challenges. Cold Spring Harb. Perspect. Med 8, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Liu D et al. Activation of mTORC1 is essential for β-adrenergic stimulation of adipose browning. J. Clin. Invest 126, 1704–1716 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Mulligan JD, Gonzalez AA, Stewart AM, Carey HV & Saupe KW Upregulation of AMPK during cold exposure occurs via distinct mechanisms in brown and white adipose tissue of the mouse. J. Physiol 580, 677–684 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Kato H et al. ER-resident sensor PERK is essential for mitochondrial thermogenesis in brown adipose tissue. Life Sci. Alliance 3, e201900576 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Lu X et al. Mitophagy controls beige adipocyte maintenance through a Parkin-dependent and UCP1-independent mechanism. Sci. Signal 11, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Reznick RM & Shulman GI The role of AMP-activated protein kinase in mitochondrial biogenesis. J. Physiol 574, 33–39 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Popov DV Adaptation of Skeletal Muscles to Contractile Activity of Varying Duration and Intensity: The Role of PGC-1α. Biochem. 83, 613–628 (2018). [DOI] [PubMed] [Google Scholar]
  • 65.Spiegelman B Hormones, Metabolisme and the Benefits of Exercise. Trends in Molecular Medicine vol. 7 (Springer International Publishing, 2001). [Google Scholar]
  • 66.Wu H et al. Regulation of mitochondrial biogenesis in skeletal muscle by caMK. Science 296, 349–352 (2002). [DOI] [PubMed] [Google Scholar]
  • 67.Gwinn DM et al. AMPK Phosphorylation of Raptor Mediates a Metabolic Checkpoint. Mol. Cell 30, 214–226 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Inoki K, Zhu T & Guan KL TSC2 Mediates Cellular Energy Response to Control Cell Growth and Survival. Cell 115, 577–590 (2003). [DOI] [PubMed] [Google Scholar]
  • 69.Sancak Y et al. The rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science 320, 1496–1501 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Sancak Y et al. Ragulator-rag complex targets mTORC1 to the lysosomal surface and is necessary for its activation by amino acids. Cell 141, 290–303 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Nandagopal N & Roux PP Regulation of global and specific mRNA translation by the mTOR signaling pathway. Translation 3, e983402 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Roux PP & Topisirovic I Regulation of mRNA translation by signaling pathways. Cold Spring Harb. Perspect. Biol 4, a012252–a012252 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Shirokikh NE & Preiss T Translation initiation by cap-dependent ribosome recruitment: Recent insights and open questions. Wiley Interdiscip. Rev. RNA 9, (2018). [DOI] [PubMed] [Google Scholar]
  • 74.Hsieh AC et al. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 485, 55–61 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Thoreen CC et al. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature 485, 109–113 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Morita M et al. MTORC1 controls mitochondrial activity and biogenesis through 4E-BP-dependent translational regulation. Cell Metab. 18, 698–711 (2013). [DOI] [PubMed] [Google Scholar]
  • 77.Larsson O et al. Distinct perturbation of the translatome by the antidiabetic drug metformin. Proc. Natl. Acad. Sci. U. S. A 109, 8977–8982 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Gandin V et al. NanoCAGE reveals 5’ UTR features that define specific modes of translation of functionally related MTOR-sensitive mRNAs. Genome Res. 26, 636–648 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Elfakess R et al. Unique translation initiation of mRNAs-containing TISU element. Nucleic Acids Res. 39, 7598–7609 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Sinvani H et al. Translational tolerance of mitochondrial genes to metabolic energy stress involves TISU and eIF1-eIF4GI cooperation in start codon selection. Cell Metab. 21, 479–492 (2015). [DOI] [PubMed] [Google Scholar]
  • 81.Elfakess R & Dikstein R A translation initiation element specific to mRNAs with very short 5′UTR that also regulates transcription. PLoS One 3, e3094 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Morita M et al. mTOR Controls Mitochondrial Dynamics and Cell Survival via MTFP1. Mol. Cell 67, 922–935.e5 (2017). [DOI] [PubMed] [Google Scholar]
  • 83.Dumesic PA et al. An Evolutionarily Conserved uORF Regulates PGC1α and Oxidative Metabolism in Mice, Flies, and Bluefin Tuna. Cell Metab. 30, 190–200.e6 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.McGillivray P et al. A comprehensive catalog of predicted functional upstream open reading frames in humans. Nucleic Acids Res. 46, 3326–3338 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Johnstone TG, Bazzini AA & Giraldez AJ Upstream ORF s are prevalent translational repressors in vertebrates. EMBO J. 35, 706–723 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Lynch MR, Tran MT & Parikh SM PGC1α in the kidney. Am. J. Physiol. - Ren. Physiol 314, F1–F8 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Han SH et al. PGC-1a protects from notch-induced kidney fibrosis development. J. Am. Soc. Nephrol 28, 3312–3322 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Pla‐Martín D et al. CLUH granules coordinate translation of mitochondrial proteins with mTORC1 signaling and mitophagy. EMBO J. 39, e102731 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Gao J et al. CLUH regulates mitochondrial biogenesis by binding mRNAs of nuclear-encoded mitochondrial proteins. J. Cell Biol 207, 213–223 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Schatton D et al. CLUH regulates mitochondrial metabolism by controlling translation and decay of target mRNAs. J. Cell Biol 216, 675–693 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Sen A & Cox RT Clueless is a conserved ribonucleoprotein that binds the ribosome at the mitochondrial outer membrane. Biol. Open 5, 195–203 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Sen A, Kalvakuri S, Bodmer R & Cox RT Clueless, a protein required for mitochondrial function, interacts with the PINK1-Parkin complex in Drosophila. DMM Dis. Model. Mech 8, 577–589 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Cannavino J et al. Regulation of cold-induced thermogenesis by the RNA binding protein FAM195A. Proc. Natl. Acad. Sci. U. S. A 118, e2104650118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Molliex A et al. Phase Separation by Low Complexity Domains Promotes Stress Granule Assembly and Drives Pathological Fibrillization. Cell 163, 123–133 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Kummer E & Ban N Mechanisms and regulation of protein synthesis in mitochondria. Nat. Rev. Mol. Cell Biol 22, 307–325 (2021). [DOI] [PubMed] [Google Scholar]
  • 96.Montoya J, Ojala D & Attardi G Distinctive features of the 5′-terminal sequences of the human mitochondrial mRNAs. Nature 290, 465–470 (1981). [DOI] [PubMed] [Google Scholar]
  • 97.Richman TR et al. Loss of the RNA-binding protein TACO1 causes late-onset mitochondrial dysfunction in mice. Nat. Commun 7, 11884 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Weraarpachai W et al. Mutation in TACO1, encoding a translational activator of COX I, results in cytochrome c oxidase deficiency and late-onset Leigh syndrome. Nat. Genet 41, 833–837 (2009). [DOI] [PubMed] [Google Scholar]
  • 99.Taggart JC & Li GW Production of Protein-Complex Components Is Stoichiometric and Lacks General Feedback Regulation in Eukaryotes. Cell Syst. 7, 580–589.e4 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Liu L et al. Nutrient sensing by the mitochondrial transcription machinery dictates oxidative phosphorylation. J. Clin. Invest 124, 768–784 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Mukaneza Y et al. MTORC1 is required for expression of LRPPRC and cytochrome-c oxidase but not HIF-1α in leigh syndrome French Canadian type patient fibroblasts. Am. J. Physiol. - Cell Physiol 317, C58–C67 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Weraarpachai W et al. Mutations in C12orf62, a factor that couples COX i synthesis with cytochrome c oxidase assembly, cause fatal neonatal lactic acidosis. Am. J. Hum. Genet 90, 142–151 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Vance JE MAM (mitochondria-associated membranes) in mammalian cells: Lipids and beyond. Biochim. Biophys. Acta - Mol. Cell Biol. Lipids 1841, 595–609 (2014). [DOI] [PubMed] [Google Scholar]
  • 104.Hornbeck PV et al. PhosphoSitePlus, 2014: Mutations, PTMs and recalibrations. Nucleic Acids Res. 43, D512–D520 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Kampjut D & Sazanov LA The coupling mechanism of mammalian respiratory complex I. Science 370, (2020). [DOI] [PubMed] [Google Scholar]
  • 106.Friedkin M & Lehninger AL Oxidation-Coupled Incorporation of Inorganic Radiophosphate Into Phospholipide and Nucleic Acid in a Cell-Free System. J. Biol. Chem 177, 775–788 (1949). [PubMed] [Google Scholar]
  • 107.Yadava N, Potluri P & Scheffler IE Investigations of the potential effects of phosphorylation of the MWFE and ESSS subunits on complex I activity and assembly. Int. J. Biochem. Cell Biol 40, 447–460 (2008). [DOI] [PubMed] [Google Scholar]
  • 108.De Rasmo D et al. Phosphorylation pattern of the NDUFS4 subunit of complex I of the mammalian respiratory chain. Mitochondrion 10, 464–471 (2010). [DOI] [PubMed] [Google Scholar]
  • 109.Samavati L, Lee I, Mathes I, Lottspeich F & Hüttemann M Tumor necrosis factor α inhibits oxidative phosphorylation through tyrosine phosphorylation at subunit I of cytochrome c oxidase. J. Biol. Chem 283, 21134–21144 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Acin-Perez R, Gatti DL, Bai Y & Manfredi G Protein phosphorylation and prevention of cytochrome oxidase inhibition by ATP: Coupled mechanisms of energy metabolism regulation. Cell Metab. 13, 712–719 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Akabane S et al. PKA Regulates PINK1 Stability and Parkin Recruitment to Damaged Mitochondria through Phosphorylation of MIC60. Mol. Cell 62, 371–384 (2016). [DOI] [PubMed] [Google Scholar]
  • 112.De Rasmo D, Panelli D, Sardanelli AM & Papa S cAMP-dependent protein kinase regulates the mitochondrial import of the nuclear encoded NDUFS4 subunit of complex I. Cell. Signal 20, 989–997 (2008). [DOI] [PubMed] [Google Scholar]
  • 113.Srinivasan S et al. Oxidative Stress Induced Mitochondrial Protein Kinase A Mediates Cytochrome C Oxidase Dysfunction. PLoS One 8, e77129 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Wang Z et al. Cyclin B1/Cdk1 coordinates mitochondrial respiration for Cell-Cycle G2/M progression. Dev. Cell 29, 217–232 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Morais VA et al. PINK1 loss-of-function mutations affect mitochondrial complex I activity via NdufA10 ubiquinone uncoupling. Science 344, 203–207 (2014). [DOI] [PubMed] [Google Scholar]
  • 116.Ogura M, Yamaki J, Homma MK & Homma Y Mitochondrial c-Src regulates cell survival through phosphorylation of respiratory chain components. Biochem. J 447, 281–289 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Salvi M, Morrice NA, Brunati AM & Toninello A Identification of the flavoprotein of succinate dehydrogenase and aconitase as in vitro mitochondrial substrates of Fgr tyrosine kinase. FEBS Lett. 581, 5579–5585 (2007). [DOI] [PubMed] [Google Scholar]
  • 118.Acín-Pérez R et al. ROS-triggered phosphorylation of complex II by Fgr kinase regulates cellular adaptation to fuel use. Cell Metab. 19, 1020–1033 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Baeza J, Smallegan MJ & Denu JM Site-specific reactivity of nonenzymatic lysine acetylation. ACS Chem. Biol 10, 122–128 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Narita T, Weinert BT & Choudhary C Functions and mechanisms of non-histone protein acetylation. Nature Reviews Molecular Cell Biology vol. 20 156–174 (2019). [DOI] [PubMed] [Google Scholar]
  • 121.Yang W et al. Mitochondrial Sirtuin Network Reveals Dynamic SIRT3-Dependent Deacetylation in Response to Membrane Depolarization. Cell 167, 985–1000.e21 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Ahn BH et al. A role for the mitochondrial deacetylase Sirt3 in regulating energy homeostasis. Proc. Natl. Acad. Sci. U. S. A 105, 14447–14452 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Porter GA, Urciuoli WR, Brookes PS & Nadtochiy SM SIRT3 deficiency exacerbates ischemia-reperfusion injury: Implication for aged hearts. Am. J. Physiol. - Hear. Circ. Physiol 306, H1602–1609 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Koentges C et al. SIRT3 deficiency impairs mitochondrial and contractile function in the heart. Basic Res. Cardiol 110, 1–20 (2015). [DOI] [PubMed] [Google Scholar]
  • 125.Parodi-Rullán RM, Chapa-Dubocq X, Rullán PJ, Jang S & Javadov S High sensitivity of SIRT3 deficient hearts to ischemia-reperfusion is associated with mitochondrial abnormalities. Front. Pharmacol 8, 275 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Finley LWS et al. Succinate dehydrogenase is a direct target of sirtuin 3 deacetylase activity. PLoS One 6, e23295 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Chen Y et al. Sirtuin-3 (SIRT3), a therapeutic target with oncogenic and tumor-suppressive function in cancer. Cell Death Dis. 5, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Xiao H et al. A Quantitative Tissue-Specific Landscape of Protein Redox Regulation during Aging. Cell 180, 968–983.e24 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Mills EL et al. Cysteine 253 of UCP1 regulates energy expenditure and sex-dependent adipose tissue inflammation. Cell Metab. 34, 140–157.e8 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Chouchani ET et al. Cardioprotection by S-nitrosation of a cysteine switch on mitochondrial complex i. Nat. Med 19, 753–759 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Yin Z et al. Structural basis for a complex I mutation that blocks pathological ROS production. Nat. Commun 12, 707.(1–12) (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Burger N et al. ND3 Cys39 in complex I is exposed during mitochondrial respiration. Cell Chem. Biol 29, 636–649.e14 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Araiso Y et al. Structure of the mitochondrial import gate reveals distinct preprotein paths. Nature 575, 395–401 (2019). [DOI] [PubMed] [Google Scholar]
  • 134.Tucker K & Park E Cryo-EM structure of the mitochondrial protein-import channel TOM complex at near-atomic resolution. Nat. Struct. Mol. Biol 26, 1158–1166 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Roise D et al. Amphiphilicity is essential for mitochondrial presequence function. EMBO J. 7, 649–653 (1988). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Allison DS & Schatz G Artificial mitochondrial presequences. Proc. Natl. Acad. Sci. U. S. A 83, 9011–9015 (1986). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Schmidt O et al. Regulation of mitochondrial protein import by cytosolic kinases. Cell 144, 227–239 (2011). [DOI] [PubMed] [Google Scholar]
  • 138.Pfanner N & Neupert W Distinct steps in the import of ADP/ATP carrier into mitochondria. J. Biol. Chem 262, 7528–7536 (1987). [PubMed] [Google Scholar]
  • 139.Pfanner N, Tropschug M & Neupert W Mitochondrial protein import: Nucleoside triphosphates are involved in conferring import-competence to precursors. Cell 49, 815–823 (1987). [DOI] [PubMed] [Google Scholar]
  • 140.Young JC, Hoogenraad NJ & Hartl FU Molecular chaperones Hsp90 and Hsp70 deliver preproteins to the mitochondrial import receptor Tom70. Cell 112, 41–50 (2003). [DOI] [PubMed] [Google Scholar]
  • 141.Bhangoo MK et al. Multiple 40-kDa heat-shock protein chaperones function in Tom70-dependent mitochondrial import. Mol. Biol. Cell 18, 3414–3428 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Rampelt H et al. The mitochondrial carrier pathway transports non-canonical substrates with an odd number of transmembrane segments. BMC Biol. 18, 2 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Yamamoto H et al. Roles of Tom70 in import of presequence-containing mitochondrial proteins. J. Biol. Chem 284, 31635–31646 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Backes S et al. Tom70 enhances mitochondrial preprotein import efficiency by binding to internal targeting sequences. J. Cell Biol 217, 1369–1382 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Fan ACY et al. Interaction between the human mitochondrial import receptors Tom20 and Tom70 in vitro suggests a chaperone displacement mechanism. J. Biol. Chem 286, 32208–32219 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Wei X et al. Mutations in TOMM70 lead to multi-OXPHOS deficiencies and cause severe anemia, lactic acidosis, and developmental delay. J. Hum. Genet 65, 231–240 (2020). [DOI] [PubMed] [Google Scholar]
  • 147.Brix J, Dietmeier K & Pfanner N Differential recognition of preproteins by the purified cytosolic domains of the mitochondrial import receptors Tom20, Tom22, and Tom70. J. Biol. Chem 272, 20730–20735 (1997). [DOI] [PubMed] [Google Scholar]
  • 148.Kreimendahl S, Schwichtenberg J, Günnewig K, Brandherm L & Rassow J The selectivity filter of the mitochondrial protein import machinery. BMC Biol. 18, 1–23 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Tripathi A, Mandon EC, Gilmore R & Rapoport TA Two alternative binding mechanisms connect the protein translocation Sec71-Sec72 complex with heat shock proteins. J. Biol. Chem 292, 8007–8018 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Gordon DE et al. Comparative host-coronavirus protein interaction networks reveal pan-viral disease mechanisms. Science 370, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Liu XY, Wei B, Shi HX, Shan YF & Wang C Tom70 mediates activation of interferon regulatory factor 3 on mitochondria. Cell Res. 20, 994–1011 (2010). [DOI] [PubMed] [Google Scholar]
  • 152.Filadi R et al. TOM70 Sustains Cell Bioenergetics by Promoting IP3R3-Mediated ER to Mitochondria Ca2+ Transfer. Curr. Biol 28, 369–382.e6 (2018). [DOI] [PubMed] [Google Scholar]
  • 153.An YA et al. Dysregulation of amyloid precursor protein impairs adipose tissue mitochondrial function and promotes obesity. Nat. Metab 1, 1243–1257 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Leney AC, El Atmioui D, Wu W, Ovaa H & Heck AJR Elucidating crosstalk mechanisms between phosphorylation and O-GlcNAcylation. Proc. Natl. Acad. Sci. U. S. A 114, E7255–E7261 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Tarrant MK et al. Regulation of CK2 by phosphorylation and O-GlcNAcylation revealed by semisynthesis. Nat. Chem. Biol 8, 262–269 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Shinoda K et al. Phosphoproteomics Identifies CK2 as a Negative Regulator of Beige Adipocyte Thermogenesis and Energy Expenditure. Cell Metab. 22, 997–1008 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Manni S et al. Protein kinase CK2 protects multiple myeloma cells from ER stress-induced apoptosis and from the cytotoxic effect of HSP90 inhibition through regulation of the unfolded protein response. Clin. Cancer Res 18, 1888–1900 (2012). [DOI] [PubMed] [Google Scholar]
  • 158.Hessenauer A, Schneider CC, Götz Claudia C & Montenarh M CK2 inhibition induces apoptosis via the ER stress response. Cell. Signal 23, 145–151 (2011). [DOI] [PubMed] [Google Scholar]
  • 159.Voos W Chaperone-protease networks in mitochondrial protein homeostasis. Biochim. Biophys. Acta - Mol. Cell Res 1833, 388–399 (2013). [DOI] [PubMed] [Google Scholar]
  • 160.Matsushima Y et al. Mitochondrial Lon protease is a gatekeeper for proteins newly imported into the matrix. Commun. Biol 4, 974 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Shin CS et al. LONP1 and mtHSP70 cooperate to promote mitochondrial protein folding. Nat. Commun 12, 265 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Rep M et al. Promotion of mitochondrial membrane complex assembly by a proteolytically inactive yeast Lon. Science 274, 103–106 (1996). [DOI] [PubMed] [Google Scholar]
  • 163.Ghosh JC et al. Akt phosphorylation of mitochondrial Lonp1 protease enables oxidative metabolism and advanced tumor traits. Oncogene 38, 6926–6939 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.MacVicar T et al. Lipid signalling drives proteolytic rewiring of mitochondria by YME1L. Nature 575, 361–365 (2019). [DOI] [PubMed] [Google Scholar]
  • 165.Pfanner N, Warscheid B & Wiedemann N Mitochondrial proteins: from biogenesis to functional networks. Nature Reviews Molecular Cell Biology vol. 20 267–284 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Priesnitz C, Pfanner N & Becker T Studying protein import into mitochondria. Methods in Cell Biology vol. 155 (Elsevier Inc., 2020). [DOI] [PubMed] [Google Scholar]
  • 167.Guerrero-Castillo S et al. The Assembly Pathway of Mitochondrial Respiratory Chain Complex I. Cell Metab. 25, 128–139 (2017). [DOI] [PubMed] [Google Scholar]
  • 168.Vercellino I & Sazanov LA The assembly, regulation and function of the mitochondrial respiratory chain. Nat. Rev. Mol. Cell Biol 23, 141–161 (2022). [DOI] [PubMed] [Google Scholar]
  • 169.Schägger H & Pfeiffer K The Ratio of Oxidative Phosphorylation Complexes I-V in Bovine Heart Mitochondria and the Composition of Respiratory Chain Supercomplexes. J. Biol. Chem 276, 37861–37867 (2001). [DOI] [PubMed] [Google Scholar]
  • 170.Greggio C et al. Enhanced Respiratory Chain Supercomplex Formation in Response to Exercise in Human Skeletal Muscle. Cell Metab. 25, 301–311 (2017). [DOI] [PubMed] [Google Scholar]
  • 171.Schägger H & Pfeiffer K Supercomplexes in the respiratory chains of yeast and mammalian mitochondria. EMBO J. 19, 1777–1783 (2000). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Acin-Perez R & Enriquez JA The function of the respiratory supercomplexes: The plasticity model. Biochimica et Biophysica Acta - Bioenergetics vol. 1837 444–450 (2014). [DOI] [PubMed] [Google Scholar]
  • 173.Acín-Pérez R, Fernández-Silva P, Peleato ML, Pérez-Martos A & Enriquez JA Respiratory Active Mitochondrial Supercomplexes. Mol. Cell 32, 529–539 (2008). [DOI] [PubMed] [Google Scholar]
  • 174.Gonzalez-Franquesa A et al. Mass-spectrometry-based proteomics reveals mitochondrial supercomplexome plasticity. Cell Rep. 35, 109180 (2021). [DOI] [PubMed] [Google Scholar]
  • 175.Lapuente-Brun E et al. Supercomplex assembly determines electron flux in the mitochondrial electron transport chain. Science 340, 1567–1570 (2013). [DOI] [PubMed] [Google Scholar]
  • 176.Lopez-Fabuel I et al. Complex I assembly into supercomplexes determines differential mitochondrial ROS production in neurons and astrocytes. Proc. Natl. Acad. Sci. U. S. A 113, 13063–13068 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Calvo E et al. Functional role of respiratory supercomplexes in mice: SCAF1 relevance and segmentation of the Qpool. Sci. Adv 6, eaba7509 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Vercellino I & Sazanov LA Structure and assembly of the mammalian mitochondrial supercomplex CIII2CIV. Nature 598, 364–367 (2021). [DOI] [PubMed] [Google Scholar]
  • 179.Moe A, Trani J Di, Rubinstein JL & Brzezinski P Cryo-EM structure and kinetics reveal electron transfer by 2D diffusion of cytochrome c in the yeast III-IV respiratory supercomplex. Proc. Natl. Acad. Sci. U. S. A 118, e2021157118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Berndtsson J et al. Respiratory supercomplexes enhance electron transport by decreasing cytochrome c diffusion distance. EMBO Rep. 21, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Blanchi C, Genova ML, Castelli GP & Lenaz G The mitochondrial respiratory chain is partially organized in a supercomplex assembly: Kinetic evidence using flux control analysis. J. Biol. Chem 279, 36562–36569 (2004). [DOI] [PubMed] [Google Scholar]
  • 182.Acín-Pérez R et al. Respiratory complex III is required to maintain complex I in mammalian mitochondria. Mol. Cell 13, 805–815 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Calvaruso MA et al. Mitochondrial complex III stabilizes complex I in the absence of NDUFS4 to provide partial activity. Hum. Mol. Genet 21, 115–120 (2012). [DOI] [PubMed] [Google Scholar]
  • 184.Schägger H et al. Significance of respirasomes for the assembly/stability of human respiratory chain complex I. J. Biol. Chem 279, 36349–36353 (2004). [DOI] [PubMed] [Google Scholar]
  • 185.Protasoni M et al. Respiratory supercomplexes act as a platform for complex III ‐mediated maturation of human mitochondrial complexes I and IV. EMBO J. 39, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Blaza JN, Serreli R, Jones AJY, Mohammed K & Hirst J Kinetic evidence against partitioning of the ubiquinone pool and the catalytic relevance of respiratory-chain supercomplexes. Proc. Natl. Acad. Sci. U. S. A 111, 15735–15740 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Lobo-Jarne T et al. Human COX7A2L Regulates Complex III Biogenesis and Promotes Supercomplex Organization Remodeling without Affecting Mitochondrial Bioenergetics. Cell Rep. 25, 1786–1799.e4 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Cogliati S et al. Mechanism of super-assembly of respiratory complexes III and IV. Nature 539, 579–582 (2016). [DOI] [PubMed] [Google Scholar]
  • 189.Pérez-Pérez R et al. COX7A2L Is a Mitochondrial Complex III Binding Protein that Stabilizes the III2+IV Supercomplex without Affecting Respirasome Formation. Cell Rep. 16, 2387–2398 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Ikeda K et al. Mitochondrial supercomplex assembly promotes breast and endometrial tumorigenesis by metabolic alterations and enhanced hypoxia tolerance. Nat. Commun 10, (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Hollinshead KER et al. Respiratory Supercomplexes Promote Mitochondrial Efficiency and Growth in Severely Hypoxic Pancreatic Cancer. Cell Rep. 33, 108231 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Mirali S et al. The mitochondrial peptidase, neurolysin, regulates respiratory chain supercomplex formation and is necessary for AML viability. Sci. Transl. Med 12, (2020). [DOI] [PubMed] [Google Scholar]
  • 193.García‐Poyatos C et al. Scaf1 promotes respiratory supercomplexes and metabolic efficiency in zebrafish. EMBO Rep. 21, e50287–e50287 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Chan DC Mitochondrial Dynamics and Its Involvement in Disease. Annu. Rev. Pathol. Mech. Dis 15, 235–259 (2020). [DOI] [PubMed] [Google Scholar]
  • 195.Friedman JR & Nunnari J Mitochondrial form and function. Nature 505, 335–343 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Friedman JR et al. ER tubules mark sites of mitochondrial division. Science 334, 358–362 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Rambold AS, Kostelecky B, Elia N & Lippincott-Schwartz J Tubular network formation protects mitochondria from autophagosomal degradation during nutrient starvation. Proc. Natl. Acad. Sci. U. S. A 108, 10190–10195 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Gomes LC, Di Benedetto G & Scorrano L During autophagy mitochondria elongate, are spared from degradation and sustain cell viability. Nat. Cell Biol 13, 589–598 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Kim H et al. Fine-Tuning of Drp1/Fis1 Availability by AKAP121/Siah2 Regulates Mitochondrial Adaptation to Hypoxia. Mol. Cell 44, 532–544 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Yao CH et al. Mitochondrial fusion supports increased oxidative phosphorylation during cell proliferation. Elife 8, (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Wikstrom JD et al. Hormone-induced mitochondrial fission is utilized by brown adipocytes as an amplification pathway for energy expenditure. EMBO J. 33, 418–436 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Coronado M et al. Physiological mitochondrial fragmentation is a normal cardiac adaptation to increased energy demand. Circ. Res 122, 282–295 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Waters LR, Ahsan FM, Wolf DM, Shirihai O & Teitell MA Initial B Cell Activation Induces Metabolic Reprogramming and Mitochondrial Remodeling. iScience 5, 99–109 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Ron-Harel N et al. Mitochondrial Biogenesis and Proteome Remodeling Promote One-Carbon Metabolism for T Cell Activation. Cell Metab. 24, 104–117 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Cogliati S et al. Mitochondrial cristae shape determines respiratory chain supercomplexes assembly and respiratory efficiency. Cell 155, 160–171 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Bal NC et al. Both brown adipose tissue and skeletal muscle thermogenesis processes are activated during mild to severe cold adaptation in mice. J. Biol. Chem 292, 16616–16625 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Kondadi AK et al. Cristae undergo continuous cycles of fusion and fission in a MICOS-dependent manner. bioRxiv. e49766, 654541 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Patten DA et al. OPA1‐dependent cristae modulation is essential for cellular adaptation to metabolic demand. EMBO J. 33, 2676–2691 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Tang J et al. Sam50–Mic19–Mic60 axis determines mitochondrial cristae architecture by mediating mitochondrial outer and inner membrane contact. Cell Death Differ. 27, 146–160 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Ott C, Dorsch E, Fraunholz M, Straub S & Kozjak-Pavlovic V Detailed analysis of the human mitochondrial contact site complex indicate a hierarchy of subunits. PLoS One 10, 1–15 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Li H et al. Mic60/Mitofilin determines MICOS assembly essential for mitochondrial dynamics and mtDNA nucleoid organization. Cell Death Differ. 23, 380–392 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Friedman JR, Mourier A, Yamada J, Michael McCaffery J & Nunnari J MICOS coordinates with respiratory complexes and lipids to establish mitochondrial inner membrane architecture. Elife 2015, 1–61 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Strauss M, Hofhaus G, Schröder RR & Kühlbrandt W Dimer ribbons of ATP synthase shape the inner mitochondrial membrane. EMBO J. 27, 1154–1160 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Daum B, Walter A, Horst A, Osiewacz HD & Kühlbrandt W Age-dependent dissociation of ATP synthase dimers and loss of inner-membrane cristae in mitochondria. Proc. Natl. Acad. Sci. U. S. A 110, 15301–15306 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Paumard P The ATP synthase is involved in generating mitochondrial cristae morphology. EMBO J. 21, 221–230 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Blum TB, Hahn A, Meier T, Davies KM & Kühlbrandt W Dimers of mitochondrial ATP synthase induce membrane curvature and self-assemble into rows. Proc. Natl. Acad. Sci. U. S. A 116, 4250–4255 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Sophie Mokas JRM et al. Uncoupling Stress Granule Assembly and Translation Initiation Inhibition. Mol. Biol. Cell 20, 2673–2683 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Kramarova TV et al. Mitochondrial ATP synthase levels in brown adipose tissue are governed by the c‐Fo subunit P1 isoform. FASEB J. 22, 55–63 (2008). [DOI] [PubMed] [Google Scholar]
  • 219.Rampelt H, Zerbes RM, van der Laan M & Pfanner N Role of the mitochondrial contact site and cristae organizing system in membrane architecture and dynamics. Biochimica et Biophysica Acta - Molecular Cell Research vol. 1864 737–746 (2017). [DOI] [PubMed] [Google Scholar]
  • 220.Banci L et al. Structural characterization of CHCHD5 and CHCHD7: Two atypical human twin CX9C proteins. J. Struct. Biol 180, 190–200 (2012). [DOI] [PubMed] [Google Scholar]
  • 221.Sakowska P et al. The Oxidation Status of Mic19 Regulates MICOS Assembly. Mol. Cell. Biol 35, 4222–4237 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Zhou ZD, Saw WT & Tan EK Mitochondrial CHCHD-Containing Proteins: Physiologic Functions and Link with Neurodegenerative Diseases. Molecular Neurobiology vol. 54 5534–5546 (2017). [DOI] [PubMed] [Google Scholar]
  • 223.Ueda E et al. Myristoyl group-aided protein import into the mitochondrial intermembrane space. Sci. Rep 9, 1185 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Tsai PI et al. PINK1 Phosphorylates MIC60/Mitofilin to Control Structural Plasticity of Mitochondrial Crista Junctions. Mol. Cell 69, 744–756.e6 (2018). [DOI] [PubMed] [Google Scholar]
  • 225.Kawano S et al. Structure-function insights into direct lipid transfer between membranes by Mmm 1-Mdm 12 of ERMES. J. Cell Biol 217, 959–974 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Guillén-Samander A et al. VPS13D bridges the ER to mitochondria and peroxisomes via Miro. J. Cell Biol 220, e202010004 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Anand R et al. MIC26 and MIC27 cooperate to regulate cardiolipin levels and the landscape of OXPHOS complexes. Life Sci. Alliance 3, 1–17 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Koob S, Barrera M, Anand R & Reichert AS The non-glycosylated isoform of MIC26 is a constituent of the mammalian MICOS complex and promotes formation of crista junctions. Biochim. Biophys. Acta - Mol. Cell Res 1853, 1551–1563 (2015). [DOI] [PubMed] [Google Scholar]
  • 229.Van Den Brink-Van Der Laan E, Antoinette Killian J & De Kruijff B Nonbilayer lipids affect peripheral and integral membrane proteins via changes in the lateral pressure profile. Biochim. Biophys. Acta - Biomembr 1666, 275–288 (2004). [DOI] [PubMed] [Google Scholar]
  • 230.Mejia EM & Hatch GM Mitochondrial phospholipids: role in mitochondrial function. J. Bioenerg. Biomembr 48, 99–112 (2016). [DOI] [PubMed] [Google Scholar]
  • 231.Osman C, Voelker DR & Langer T Making heads or tails of phospholipids in mitochondria. J. Cell Biol 192, 7–16 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Pfeiffer K et al. Cardiolipin Stabilizes Respiratory Chain Supercomplexes. J. Biol. Chem 278, 52873–52880 (2003). [DOI] [PubMed] [Google Scholar]
  • 233.Böttinger L et al. Phosphatidylethanolamine and cardiolipin differentially affect the stability of mitochondrial respiratory chain supercomplexes. J. Mol. Biol 423, 677–686 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Tasseva G et al. Phosphatidylethanolamine deficiency in mammalian mitochondria impairs oxidative phosphorylation and alters mitochondrial morphology. J. Biol. Chem 288, 4158–4173 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Das S et al. ATP citrate lyase improves mitochondrial function in skeletal muscle. Cell Metab. 21, 868–876 (2015). [DOI] [PubMed] [Google Scholar]
  • 236.Lynes MD et al. Cold-Activated Lipid Dynamics in Adipose Tissue Highlights a Role for Cardiolipin in Thermogenic Metabolism. Cell Rep. 24, 781–790 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.May FJ et al. Lipidomic Adaptations in White and Brown Adipose Tissue in Response to Exercise Demonstrate Molecular Species-Specific Remodeling. Cell Rep. 18, 1558–1572 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Marcher AB et al. RNA-Seq and Mass-Spectrometry-Based Lipidomics Reveal Extensive Changes of Glycerolipid Pathways in Brown Adipose Tissue in Response to Cold. Cell Rep. 13, 2000–2013 (2015). [DOI] [PubMed] [Google Scholar]
  • 239.Jain IH et al. Genetic Screen for Cell Fitness in High or Low Oxygen Highlights Mitochondrial and Lipid Metabolism. Cell 181, 716–727.e11 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Baker CD, Ball WB, Pryce EN & Gohil VM Specific requirements of nonbilayer phospholipids in mitochondrial respiratory chain function and formation. Mol. Biol. Cell 27, 2161–2171 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Tuller G, Nemec T, Hrastnik C & Daum G Lipid composition of subcellular membranes of an FY1679-derived haploid yeast wild-type strain grown on different carbon sources. Yeast 15, 1555–1564 (1999). [DOI] [PubMed] [Google Scholar]
  • 242.Yan F, Zhao H & Zeng Y Lipidomics: a promising cancer biomarker. Clin. Transl. Med 7, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Dean JM & Lodhi IJ Structural and functional roles of ether lipids. Protein Cell 9, 196–206 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Park H et al. Peroxisome-derived lipids regulate adipose thermogenesis by mediating cold-induced mitochondrial fission. J. Clin. Invest 129, 694–711 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Kimura T et al. Substantial Decrease in Plasmalogen in the Heart Associated with Tafazzin Deficiency. Biochemistry 57, 2162–2175 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Kimura T et al. Plasmalogen loss caused by remodeling deficiency in mitochondria. Life Sci. Alliance 2, e201900348 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Zhu Y et al. Alkylglyceronephosphate synthase (AGPS) alters lipid signaling pathways and supports chemotherapy resistance of glioma and hepatic carcinoma cell lines. Asian Pacific J. Cancer Prev 15, 3219–3226 (2014). [DOI] [PubMed] [Google Scholar]
  • 248.Benjamin DI et al. Ether lipid generating enzyme AGPS alters the balance of structural and signaling lipids to fuel cancer pathogenicity. Proc. Natl. Acad. Sci. U. S. A 110, 14912–14917 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Messias MCF, Mecatti GC, Priolli DG & De Oliveira Carvalho P Plasmalogen lipids: Functional mechanism and their involvement in gastrointestinal cancer. Lipids Health Dis. 17, 41 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Wu M, Gu J, Guo R, Huang Y & Yang M Structure of Mammalian Respiratory Supercomplex I1III2IV1. Cell 167, 1598–1609.e10 (2016). [DOI] [PubMed] [Google Scholar]
  • 251.Sun F et al. Crystal Structure of Mitochondrial Respiratory Membrane Protein Complex II. Cell 121, 1043–1057 (2005). [DOI] [PubMed] [Google Scholar]
  • 252.Spikes TE, Montgomery MG & Walker JE Structure of the dimeric ATP synthase from bovine mitochondria. Proc. Natl. Acad. Sci 117, 23519–23526 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Pebay-Peyroula E et al. Structure of mitochondrial ADP/ATP carrier in complex with carboxyatractyloside. Nature 426, 39–44 (2003). [DOI] [PubMed] [Google Scholar]
  • 254.Berardi MJ, Shih WM, Harrison SC & Chou JJ Mitochondrial uncoupling protein 2 structure determined by NMR molecular fragment searching. Nature 476, 109–113 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Park Y, Reyna-Neyra A, Philippe L & Thoreen CC mTORC1 Balances Cellular Amino Acid Supply with Demand for Protein Synthesis through Post-transcriptional Control of ATF4. Cell Rep. 19, 1083–1090 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Davies KM, Anselmi C, Wittig I, Faraldo-Gómez JD & Kühlbrandt W Structure of the yeast F 1F o-ATP synthase dimer and its role in shaping the mitochondrial cristae. Proc. Natl. Acad. Sci. U. S. A 109, 13602–13607 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Nicholls DG & Ferguson SJ Respiratory Chains. in Bioenergetics 91–157 (Elsevier, 2013). doi: 10.1016/b978-0-12-388425-1.00005-1. [DOI] [Google Scholar]
  • 258.Guo R, Gu J, Zong S, Wu M & Yang M Structure and mechanism of mitochondrial electron transport chain. Biomed. J 41, 9–20 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Letts JA & Sazanov LA Clarifying the supercomplex: The higher-order organization of the mitochondrial electron transport chain. Nat. Struct. Mol. Biol 24, 800–808 (2017). [DOI] [PubMed] [Google Scholar]
  • 260.Sousa JS, Mills DJ, Vonck J & Kühlbrandt W Functional asymmetry and electron flow in the bovine respirasome. Elife 5, (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Gu J et al. The architecture of the mammalian respirasome. Nature 537, 639–643 (2016). [DOI] [PubMed] [Google Scholar]
  • 262.Letts JA, Fiedorczuk K & Sazanov LA The architecture of respiratory supercomplexes. Nature 537, 644–648 (2016). [DOI] [PubMed] [Google Scholar]
  • 263.Martínez-Reyes I & Chandel NS Mitochondrial TCA cycle metabolites control physiology and disease. Nat. Commun 11, 102 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Arnold S & Kadenbach B Cell respiration is controlled by ATP, an allosteric inhibitor of cytochrome-c oxidase. Eur. J. Biochem 249, 350–354 (1997). [DOI] [PubMed] [Google Scholar]
  • 265.Schneeberger M et al. XMitofusin 2 in POMC neurons connects ER stress with leptin resistance and energy imbalance. Cell 155, (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Tubbs E et al. Mitochondria-associated endoplasmic reticulum membrane (MAM) integrity is required for insulin signaling and is implicated in hepatic insulin resistance. Diabetes 63, 3279–3294 (2014). [DOI] [PubMed] [Google Scholar]
  • 267.De Brito OM & Scorrano L Mitofusin 2 tethers endoplasmic reticulum to mitochondria. Nature 456, 605–610 (2008). [DOI] [PubMed] [Google Scholar]
  • 268.Muñoz JP et al. Mfn2 modulates the UPR and mitochondrial function via repression of PERK. EMBO J. 32, 2348–2361 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Abrisch RG, Gumbin SC, Wisniewski BT, Lackner LL & Voeltz GK Fission and fusion machineries converge at ER contact sites to regulate mitochondrial morphology. J. Cell Biol 219, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Bravo R et al. Increased ER-mitochondrial coupling promotes mitochondrial respiration and bioenergetics during early phases of ER stress. Journal of Cell Science vol. 124 2143–2152 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Basso V, Marchesan E & Ziviani E A trio has turned into a quartet: DJ-1 interacts with the IP3R-Grp75-VDAC complex to control ER-mitochondria interaction. Cell Calcium vol. 87 (2020). [DOI] [PubMed] [Google Scholar]
  • 272.De vos KJ et al. VAPB interacts with the mitochondrial protein PTPIP51 to regulate calcium homeostasis. Hum. Mol. Genet 21, 1299–1311 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Stoica R et al. ER-mitochondria associations are regulated by the VAPB-PTPIP51 interaction and are disrupted by ALS/FTD-associated TDP-43. Nat. Commun 5, 3996 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Stoica R et al. ALS / FTD ‐associated FUS activates GSK ‐3β to disrupt the VAPB – PTPIP 51 interaction and ER –mitochondria associations. EMBO Rep. 17, 1326–1342 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Naon D et al. Critical reappraisal confirms that Mitofusin 2 is an endoplasmic reticulum-mitochondria tether. Proc. Natl. Acad. Sci. U. S. A 113, 11249–11254 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Mancuso M et al. Fatigue and exercise intolerance in mitochondrial diseases. Literature revision and experience of the Italian Network of mitochondrial diseases. Neuromuscul. Disord 22, S226–S229 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Mito T et al. Mosaic dysfunction of mitophagy in mitochondrial muscle disease. Cell Metab. 34, 197–208.e5 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Kleiner S et al. Development of insulin resistance in mice lacking PGC-1α in adipose tissues. Proc. Natl. Acad. Sci. U. S. A 109, 9635–9640 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Cohen P et al. Ablation of PRDM16 and beige adipose causes metabolic dysfunction and a subcutaneous to visceral fat switch. Cell 156, 304–316 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.Lehman JJ & Kelly DP Transcriptional activation of energy metabolic switches in the developing and hypertrophied heart. Clin. Exp. Pharmacol. Physiol 29, 339–345 (2002). [DOI] [PubMed] [Google Scholar]
  • 281.Chambers JM & Wingert RA PGC-1α in Disease: Recent Renal Insights into a Versatile Metabolic Regulator. Cells 9, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Zhang L, Liu J, Zhou F, Wang W & Chen N PGC-1α ameliorates kidney fibrosis in mice with diabetic kidney disease through an antioxidative mechanism. Mol. Med. Rep 17, 4490–4498 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Piccinin E et al. PGC-1s in the spotlight with Parkinson’s disease. Int. J. Mol. Sci 22, (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Kumar M et al. Defects in Mitochondrial Biogenesis Drive Mitochondrial Alterations in PARKIN-Deficient Human Dopamine Neurons. Stem Cell Reports 15, 629–645 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 285.Lv J et al. PGC-1α sparks the fire of neuroprotection against neurodegenerative disorders. Ageing Res. Rev 44, 8–21 (2018). [DOI] [PubMed] [Google Scholar]
  • 286.Dumauthioz N et al. Enforced PGC-1α expression promotes CD8 T cell fitness, memory formation and antitumor immunity. Cell. Mol. Immunol 18, 1761–1771 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Gerbec ZJ et al. Conditional Deletion of PGC-1α Results in Energetic and Functional Defects in NK Cells. iScience 23, 101454 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Bertero E, Kutschka I, Maack C & Dudek J Cardiolipin remodeling in Barth syndrome and other hereditary cardiomyopathies. Biochim. Biophys. Acta - Mol. Basis Dis 1866, 165803 (2020). [DOI] [PubMed] [Google Scholar]
  • 289.MacKenzie JA & Payne RM Mitochondrial protein import and human health and disease. Biochim. Biophys. Acta - Mol. Basis Dis 1772, 509–523 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES