Mitochondrial respiratory chain composition and organization in response to changing oxygen levels
Alba Timón-Gómez and Antoni Barrientos
Author Information

1Department of Neurology, University of Miami Miller School of Medicine,Miami, FL 33136

2Department of Neurology and Department of Biochemistry and MolecularBiology. University of Miami MillerSchool of Medicine, Miami, FL 33136.

*Correspondence: axt809@med.miami.edu


Citation Information
Timón-Gómez and Barrientos, J Life Sci, Vol. 2, No. 2, June 2020:1-17 doi.org/10.36069/JoLS/20200601 PMID: 32551463, PMCID: PMC7302114

Abstract

Mitochondria are the major consumer of oxygen in eukaryotic cells, owing to the requirement of oxygen to generate ATP through the mitochondrial respiratory chain (MRC) and the oxidative phosphorylation system (OXPHOS). This aerobic energy transduction is more efficient than anaerobic processes such as glycolysis. Hypoxia, a condition in which environmental or intracellular oxygen levels are below the standard range, triggers an adaptive signaling pathway within the cell. When oxygen concentrations are low, hypoxia-inducible factors (HIFs) become stabilized and activated to mount a transcriptional response that triggers modulation of cellular metabolism to adjust to hypoxic conditions. Mitochondrial aerobic metabolism is one of the main targets of the hypoxic response to regulate its functioning and efficiency in the presence of decreased oxygen levels. During evolution, eukaryotic cells and tissues have increased the plasticity of their mitochondrial OXPHOS system to cope with metabolic needs in different oxygen contexts. In mammalian mitochondria, two factors contribute to this plasticity. First, several subunits of the multimeric MRC complexes I and IV exist in multiple tissue-specific and condition-specific isoforms. Second, the MRC enzymes can coexist organized as individual entities or forming supramolecular structures known as supercomplexes, perhaps in a dynamic manner to respond to environmental conditions and cellular metabolic demands. In this review, we will summarize the information currently available on oxygen-related changes in MRC composition and organization and will discuss gaps of knowledge and research opportunities in the field.

Keywords: Hypoxia, Mitochondrial OXPHOS, MRC, hypoxia-inducible factors (HIFs)

1. Introduction

Eukaryotic cells have developed complex molecular signaling pathways to maintain homeostasis upon changing environmental and nutritional conditions. Oxygen is a crucial determinant for cellular aerobic energy transduction, and its deficit is linked to the pathogenesis of severe human diseases, such as stroke, myocardial infarction, and chronic obstructive pulmonary disease. Therefore, cells present multiple oxygen sensing and response pathways to minimize cellular damage caused by hypoxia; that is, a decrease in the availability of oxygen. The main regulators of the rapid and efficient response to hypoxia are a family of transcription factors known as hypoxia-inducible factors (HIFs). HIFs are heterodimeric factors, composed of an oxygen-regulated alpha subunit (HIF-α) and a constitutively-expressed beta subunit (HIF-β). The alpha subunit is synthesized continuously in the cytosol and degraded by the proteasome in the presence of oxygen, whereas the beta subunit is proteolytically stable (1-3). As the oxygen levels drop (threshold levels are tissue-specific (4-6), the alpha subunit is stabilized, and the HIF complex is formed and translocated to the nucleus of the cell, where it binds to conserved regions, named hypoxia-responsive elements (HRE), in the promoter or the enhancer sequences of the hypoxia-regulated genes 2. In this way, hypoxia-responsive genes are transcriptionally activated to adapt to low oxygen conditions. Many genes activated in a HIF-dependent manner belong to functional clusters involved in oxygen homeostasis modulation and metabolic regulation (7, 8). Simultaneously, a set of genes, mostly involved in cell proliferation 9, undergo an indirect decrease in their expression by transactivation of genes encoding chromatin-modifying enzymes (10,11), transcriptional repressors (12,13), or microRNAs (14).

Mitochondria are the major oxygen-consuming organelles of the cell because it houses the oxidative phosphorylation (OXPHOS) reaction, the process that couples oxygen consumption by the mitochondrial respiratory chain (MRC) with energy transduction to the chemical form of adenosine triphosphate (ATP). Upon oxygen deprivation, anaerobic energy transduction needs to take control to maintain the minimum cellular needs and, therefore, there are major molecular changes to shift from oxidative mitochondrial metabolism to glycolysis (15,16). The cellular adaptation to hypoxia also involves a decrease in energy demand by inhibition of processes, such as general protein synthesis and cell division (17). Intracellularly, mitochondria undergo distinct modifications in their morphology and distribution to adapt their function to low oxygen levels (reviewed in (18,19). These changes include a decrease of mitochondrial mass, both by suppression of mitochondrial biogenesis (20,21) and induction of mitophagy (22); changes in intracellular mitochondrial distribution towards a perinuclear accumulation (23); and modifications of mitochondrial morphology, promoting either fission in acute hypoxia (24), or fusion under chronic hypoxia to protect from apoptosis (25).

In this manuscript, we aim to briefly review the current literature on hypoxia-induced adjustments in MRC biogenesis and organization, and their biological significance. Specifically, we will discuss the switch of several MRC complex subunits to hypoxic isoforms, and modifications in the organization of MRC complexes into individual entities or assembled as macromolecular enzymes called supercomplexes (SCs), when cells and tissues are exposed to low oxygen levels.                

2. Oxygen utilization by the mitochondrial respiratory chain

The MRC facilitates electron transfer from reducing equivalents (NADH and FADH2) to molecular oxygen, a process coupled to the generation of a proton gradient across the mitochondrial inner membrane that is used by the F1Fo-ATP synthase, to catalyze the phosphorylation of ADP to ATP. The MRC is composed of four multiprotein enzymatic complexes (CI to CIV) and two mobile electron carriers (ubiquinone and cytochrome c). CI, CIII, and CIV are responsible for the proton-motive gradient between the mitochondrial matrix and the intermembrane space. The MRC CIV, or cytochrome c oxidase (COX), is the terminal oxidase of the pathway, to whose catalytic center molecular oxygen binds and is sequentially reduced to H2O. Consequently, CIV presents a high affinity for oxygen (26) and is considered the major cellular consumer of this molecule. During the MRC functioning, electrons can escape the pathway prematurely and contribute to the generation of reactive oxygen species (ROS) (27). Although, in most cases, small-molecule electron carriers such as NADH or CoQH2 (reduced coenzyme Q) do not react with O2 to generate superoxide anion O2-, it can take place at redox-active prosthetic groups within MRC proteins, or when electron carriers such as CoQH2 are bound to proteins (28). Specifically, CI and CIII are considered main sites of ROS production in the MRC (28). ROS can cause severe damage to lipids, proteins, and DNA within the cell. Thus, mitochondria contain several antioxidant defense systems to protect from those free radicals that are generated during the oxidative metabolism, such as detoxifying enzymes or glutathione. Excessive ROS production by mammalian mitochondria underlies oxidative damage in an array of human pathologies (29-33). However, physiological ROS generation contributes to retrograde redox signaling from the organelle to the cytosol and nucleus and may play an important role in the adaptation to an array of stressors, including hypoxia (34-38).

The MRC enzymatic complexes are located in the inner mitochondrial membrane; however, their organization has been a matter of intense debate over the last 50 years. Two opposed models have been historically proposed to explain the MRC organization. Whereas the solid-state model poses that MRC complexes are organized in a single rigid macromolecular assembly, the fluid-state model views the individual complexes freely diffusing in the mitochondrial inner membrane. Based on extensive data, the existence of mitochondrial supercomplexes (SCs) is today accepted and the prevalent model for the MRC organization is known as the plasticity model (39), in which supercomplexes of variable composition dynamically coexist with individual complexes presumable to facilitate fast adaptation to changes in cellular metabolism (40,41). The composition and abundance of these SCs vary depending on the cell type, tissue, organism, and even the environmental and nutritional state of the cell. In human cells, apart from CIII and CIV associations (SC CIII2-CIV), CI can bind to CIII2 (SC CI-CIII2), and together to CIV, forming the so-called respirasome (CI-CIII2-CIV1-n). Recently, cryo-EM structures of mammalian SC CI-CIII2 (42), mammalian respirasome (43-46), and human megacomplex (47) have been obtained, confirming their existence and facilitating the study of their function. The MRC organization into supercomplexes has provided a rationale for the fact that mutations directly affecting a single complex of the MRC frequently result in multienzymatic deficiencies in patients suffering from mitochondrial diseases. Besides, the disruption of these SCs is associated with a large number of mitochondrial disorders and age-associated human diseases (48-50). However, despite the physiological and biomedical relevance of this MRC organization, little is known about the regulation of SC dynamics, the factors involved, and the functional benefits that the SC organization might provide.               

There are several hypotheses about the functional advantages of the MRC organization in SCs (reviewed in (51-53)), including substrate channeling, decreased ROS production, stabilization of CI, regulation of MRC activity, and/or prevention of protein aggregation in the inner membrane. Despite numerous investigations, until now, only the hypothesis of quinone channeling as a kinetic advantage of respiratory SCs has been discarded based on experiments using an alternative quinol oxidase (54) and evidence provided by structural approaches (42). A recent theoretical modeling approach, however, has indicated that forming respiratory supercomplexes may provide a kinetic advantage by linking complexes III and IV. This model has shown that the electron flux through these complexes can be limited by diffusion of cytochrome c, and, therefore, minimizing the distance between these complexes is kinetically advantageous (55).

In support of the relevance of SCs for CI assembly and stabilization, it was recently shown in a Ndufs4-KO mouse model (56) and in a homoplasmic MTCYB-deficient human cell line, that the incorporation of the NADH module to CI was deficient in the absence of CIII (57). Several lines of evidence also reinforce the ROS hypothesis. Maranzana and colleagues also provided the first direct demonstration that the disruption of SC I+III2 leads to an increase in superoxide generation from CI (58). Additional observations have linked SC dissociation with higher ROS production: (i) lymphoblasts from Barth syndrome patients, biochemically characterized by reduced concentration and altered composition of the mitochondrial lipid cardiolipin, which leads to destabilized SCs, showed higher production of superoxide compared to control cells (59,60); (ii) a decreased stability of SCs in a mouse model lacking catalytic CIII subunit UQCRFS1 (or RISP) was associated with increased ROS levels (61); (iii) K-RAS transformed fibroblasts presented a lesser amount of SCs, correlated with higher ROS production (62, 63); and (iv) knockdown of the CI subunit NDUFS1 in neurons led to decreased incorporation of CI into SCs, impaired oxygen consumption, and increased ROS levels (49). Nevertheless, several of these hypotheses may be simultaneously correct, even if in a tissue- or condition-specific context and, therefore, further investigations are needed to establish the validity of these hypotheses and the functional relevance of MRC organization into SCs.

3. Mitochondrial respiratory chain and hypoxia

3.1. Changes in the composition of MRC complexes

When oxygen levels decrease, the activation of HIF signaling to enhance anaerobic ATP production is accompanied by a simultaneous downregulation of MRC biogenesis. However, during the transition to these adaptations, at the physiological level, a slower MRC electron transfer activity in the presence of decreasing oxygen levels also increases the chances of ROS generation, especially the superoxide anion O2- (64,65). Stimulation of ROS production under hypoxia mostly occurs at the CI, CII, and CIII sites (66,67), which concurrently activates several antioxidant defense pathways within the cell. An adaptation to minimize ROS production involves changes in the expression of tailored isoforms of several MRC complex subunits. The most targeted complexes for these adaptations are CI, a dominant acceptor of electrons, and CIV, the terminal MRC oxidase that regulates the overall electron flow.

CI undergoes several changes to decrease its activity. Hypoxia provokes a CI conformational change from an active to a dormant form, which prevents a burst of ROS generation following reoxygenation (68, 69). The dormant form is a Na+/H+ antiporter, independent of energy transduction, and formed spontaneously by a lack of substrates (70). Under chronic hypoxia, HIF-1α induces the degradation of the transmembrane CI assembly factor TMEM126B, which attenuates functional complex I assembly and reduces the cellular respiratory capacity (71). Moreover, a hypoxia-responsive microRNA, miR-210, attenuates the expression of iron-sulfur cluster assembly proteins 1 and 2 (ISCU1/2) (72-74), affecting the maturation of mitochondrial iron-sulfur containing proteins, such as NDUFS in CI and SDHD in CII (75), and hence affecting MRC complex assembly.

CIV is also modified to respond to changing oxygen levels by undergoing post-translational modification of its subunits (76-78), or by changing its composition (79, 80), to regulate energy production. Mammalian CIV contains a COX4 subunit, which exists in a pair of normoxic/hypoxic isoforms: COX4-1 and COX4-2. COX4-2 is expressed under hypoxic conditions (81), and its presence enhances the CIV catalytic constant, thereby increasing overall MRC efficiency (82). Concurrently, under low oxygen levels conditions, the Lon protease is upregulated by HIF-1 and degrades the COX4-1 subunit (83). By regulating the proportion of each isoform, cells can adjust the MRC electron transfer rate to oxygen availability, thereby increasing the efficiency of electron transfer while minimizing ROS generation when oxygen availability is reduced. Paradoxically, mammalian COX4-2 is mostly expressed in the lung, a tissue exposed to the highest concentration of oxygen. By ensuring a higher CIV efficiency, the presence of COX4-2 could act as a mechanism of protection and energetic adaptation to minimize ROS in highly oxygenated tissues (81, 84, 85). A Hypoxia-Inducible Gene Domain family (HIGD) protein, HIGD1A, was similarly described to be induced under hypoxia and to bind CIV around its heme a active center to ensure optimal CIV activity, exerting a protective function under hypoxia (86). In addition, the CIV subunit NDUFA4 is a target of miR-210 and is decreased in a HIF-dependent manner (87). The decrease of NDUFA4 coincides with the increase of its hypoxic isoform NDUFA4L2, which lowers CI and CIV activities and, thus, mitochondrial respiration, preventing excessive ROS production and maintaining the mitochondrial membrane potential (88). NDUFA4L2 expression was recently correlated with poor patient survival in hepatocellular carcinoma (89). Finally, the CIV assembly factor COX10, also a target of miR-210, and the CIV subunit COX5B (90) are decreased in a HIF-dependent manner, attenuating CIV assembly under low oxygen levels.

In summary, cells undergo a decrease in the abundance of MRC CI and CIV (see Figure 1), whereas the expression of hypoxic subunit isoforms from these complexes are activated to decrease ROS production and optimize electron transfer and energy production under hypoxic conditions.

3.2. Changes in the MRC organization into supercomplexes

The plasticity model of the MRC organization postulates that individual respiratory complexes dynamically associate into SCs to adapt to changing environmental and nutritional conditions (91). Several studies have shown that SCs are increased in human (92) and rat (93) mitochondria after exercise, participating in the antioxidant effect of physical activity. Similarly, ER stress also increases the rate of SC formation in mitochondria in human cell lines (94). On  the contrary, after starvation (39), during aging (95), or in multiple human diseases (96), the abundance of respiratory SCs is decreased. All these studies point towards a link between MRC supercomplex formation and the metabolic and/or nutritional state of the cell. However, these investigations are based on the steady-state levels of MRC complexes and SCs analyzed by Blue Native PAGE (BN-PAGE), which cannot provide information regarding the capacity of MRC complexes to dynamically associate and dissociate. There have been several attempts to overcome these methodological challenges, as reviewed in (97). The development of proximity-dependent labeling techniques (e.g., BioID) has allowed identifying transient or weak protein-protein interactions, even in the case of insoluble complexes (98). Crosslinking mass spectrometry also supported the existence of SCs in intact tissue or mitochondria (99-101). Both techniques may reflect the organization of the proteins into SCs in a more native manner than BN-PAGE because there is no extraction of the MRC complexes from the mitochondrial inner membrane. But again, their results only provide information in a steady-state manner. The recent use of GFP-FRET technique to determine SCs formation in live cells (102) has better prospects regarding dynamism information. The authors analyzed in this study fluorescent sensor proteins located in the MRC SCs by fluorescence lifetime imaging microscopy (FLIM) to monitor SC assembly and plasticity in live cells. The use of this technique could be useful to establish relationships between MRC organization and cellular nutritional and environmental conditions.

Nevertheless, there is a particular paucity of investigations on the rearrangements that the mitochondrial SCs might undergo under low oxygen levels conditions. In plants, CI becomes destabilized and dissociates from SCs after prolonged hypoxic conditions 103 (Figure 1). It was proposed that this disruption of CI-containing SCs is due to a transition from an active to a dormant form of CI, alongside with activation of alternative oxidases that could prevent the inhibition of the glycolytic pathway because of a deficiency of NAD+. A different scenario was described in freshwater turtles Trachemys scripta, known to tolerate severe hypoxia and reoxygenation without suffering from heart damage. Heart mitochondria from these turtles contain highly stable CI-containing SCs, which remain intact even in the presence of the detergent dodecyl maltoside, contrary to what is observed in mammalian SCs from other species (104). Although ROS generation was found attenuated in heart mitochondria from anoxia-acclimated turtles, the authors could not establish a direct relationship between SC formation or stability and protection against hypoxia.

 

Figure 1. Schematic representation of known modifications in the mitochondrial respiratory chain composition and organization during hypoxia.

Changes induced in complexes and supercomplexes (SCs) of the mitochondrial respiratory chain under low oxygen levels. CI undergoes a switch from its active to its dormant form, which has been related to destabilization of CI in the SCs in plants. HIF1 induces the expression of hypoxic isoforms of subunits of CIV NDUFA4L2 (SWISS-MODEL Q9NRX3) and COX4-2, while upregulating LON protease that degrades COX4-1 isoform of CIV. There is also a HIF1-dependent induction of HIGD1A (RCSB PDB 2LOM), to generate a more efficient CIV, and a decrease of the CIV subunit COX5B. Concurrently, miR-210 is activated and downregulates ISCU1/2, decreasing the levels of NDUFS subunits of CI and SDHD of CII. CIV subunit NDUFA4 and CIV assembly factor COX10 are also targets of miR-210, lowering their levels.

CI, complex I (dormant state RCSB PDB 5O31; active state RCSB PDB 5XTD). CII, avian complex II (RCSB PDB 1YQ3). CIII, complex III (RCSB PDB 5XTE). CIV, complex IV (RCSB PDB 5X62). SC, respiratory supercomplex I+III2+IV or respirasome (RCSB PDB 5GPN). HIF, Hypoxia Inducible Factor. HRE, Hypoxia Responsive Element. IMS, Intermembrane Space.

The single mammalian SC assembly factor described so far, COX7A2L (COX7RP, or SCAFI), is involved in the assembly of CIII2-CIV and in higher-order structures (CI-CIII2-CIV2-n) (39, 105-107). COX7A2L was discovered in mice (named COX7RP) simultaneously by two different groups. Lapuente-Brun and colleagues performed a screening for proteins present in SCs but not in free complexes, showing that COX7A2L promotes interactions between CIII2 and CIV (39). However, it is now accepted that COX7A2L not only associates with SCs but also independently interacts with both CIII2 and free CIV to promote SC III2+IV1 stabilization, without affecting the formation of the respirasome or SC I+III2+IV1. Lapuente-Brun and colleagues also reported that some wild-type mouse strains (e.g., C57BL/6J and BALB/c) express only a short, unstable COX7A2L isoform that failed to support CIV association into SCs, thereby promoting differences in mitochondrial respiration rates and ATP production (39). However, although the degree of respirasome instability varies among tissues (108), the functional deficits originally claimed remain under debate (51). Concurrently, Ikeda and co-workers generated a COX7A2L-KO mouse, after the identification of this protein as estrogen-sensitive (109) with a homologous sequence to COX7A subunit; and described an impairment in SC formation in muscle (110).

Expression of COX7A2L was shown to be upregulated in a tissue-specific manner in human cell lines during heat-shock and acute oxidative stress; however, the SC organization was not altered in these conditions (105). Levels of COX7A2L were also increased after carbon source switch from glucose to galactose, which stimulates mitochondrial energy metabolism (111), in several cell lines (94, 105). However, Lobo-Jarne and colleagues did not find any evident impact in MRC biogenesis and mitochondrial bioenergetics in the absence of COX7A2L in HEK293T or U87 cells (105), whereas Balsa and colleagues described a decrease in SC levels and oxygen consumption in human U2OS cells and mouse fibroblasts lacking COX7A2L in the presence of galactose (94). Its homolog protein in mice was shown to play a role in glucose metabolism (112). These differences in the results could be due to a tissue- or species-specificity of SC rearrangements or to technical issues; e.g., varying experimental conditions (use of isolated clonal KO cell lines vs. pools of clones, whole cell vs intact mitochondria analysis, or membrane solubilization conditions) and/or methods (polarography or high-resolution respirometry vs. Seahorse Bioanalyzer measurements) between groups. Therefore, the role of COX7A2L under nutritional stress remains to be fully resolved.

Recently, however, overexpression of COX7A2L in human cells was described to promote cell growth during hypoxia, inhibiting the hypoxia-induced generation of mitochondrial ROS (113). Furthermore, the authors observed stimulation of SCs CI-CIII-CIVn and CIII2-CIV2 assembly and stabilization during hypoxia when COX7A2L was overexpressed, together with an increase in the maximum respiratory rate of the cells (113). This could indicate a possible function of mitochondrial respiratory SCs in the adaptation to different environmental conditions, conceivably by enhancing respiratory efficiency while decreasing ROS production. However, more investigations are necessary to study SC biogenesis and organization under hypoxic conditions, in order to understand the possible role of these macrostructures and their significance in the adaptation to hypoxia.

The existence of other SC assembly factors involved in the adaptation of the MRC to different conditions is also plausible. The HIGD1A and HIGD2A proteins, homologous of the yeast Respiratory superComplex Factor Rcf1, are good candidates to perform a role in SC assembly and regulation during hypoxia. Rcf1 was proposed to have a role in the assembly and stability of yeast SCs (114, 115). Rcf1 and its homologous HIGD proteins belong to the Hypoxia Inducible Gene Domain 1 family and, therefore, are overexpressed under hypoxic conditions, increasing cell survival (86, 116-117). In hypoxia, HIGD1A was shown to interact with CIII and CIV (86, 116), and in standard cell culture conditions, complexome analysis showed HIGD1A to associate with SCs (118). HIGD2A was able to promote CIV and CIV-containing SCs formation in a yeast deletion model of Rcf1 (114). Recently, our group created HEK293T HIGD-KO cell lines to study the role of HIGD proteins under standard cell culture conditions. HIGD2A was described as a CIV-assembly factor, controlling and coordinating modular assembly of isolated and supercomplexed CIV (119). Differently, HIGD1A has a function in the incorporation of the CIII catalytic subunit UQCRFS1, regulating the kinetics of CIII and CIII-containing SCs. HIGD1A also binds to CIV subunits COX4-1 and COX5A and, when overexpressed, is able to suppress the CIV biogenetic and respiratory defect of HIGD2A-KO cells. Therefore, we concluded that both HIGD proteins play independent and overlapping roles in the biogenesis of respiratory complexes and SCs in physiological conditions (119). However, the mechanism of action and the role in MRC assembly and organization of these HIGD proteins in hypoxic conditions remain to be investigated.

4. Concluding remarks

Oxygen is a critical molecular component of the mitochondrial energetic metabolism, and eukaryotic cells have evolved to adjust to different levels of this molecule to survive. HIF complexes are responsible for orchestrating the cellular response to low oxygen levels, activating and suppressing the expression of diverse genes. Mitochondria are one of the main targets in the hypoxic adaptive response. Recently, multiple studies are showing that remodeling of the MRC organization into SCs is part of the adaptation of cellular metabolism to changing environmental and nutritional conditions. Exposure to hypoxia has been proposed as a candidate therapeutic option to treat mitochondrial diseases. Several studies in mice and cellular models of Leigh syndrome, an encephalomyopathy associated with MRC malfunction, showed that hypoxia attenuated the symptoms and increased the life span of the disease models. Accordingly, the authors of this review speculate that hypoxic conditions could induce modifications in the MRC, not only at the composition level but also at the organizational level. The relationship between hypoxia and human diseases accentuates the interest to explore the molecular mechanisms that modulate the MRC upon low oxygen levels, as a way to increase our knowledge of the pathogenesis of mitochondrial diseases.

Several technical and conceptual open questions remaining, some of which are listed here, are a priority for future work. For example, an optimized and standardized method to evaluate MRC function at different oxygen tensions is necessary to provide consistent results within the scientific community. For this purpose, high-resolution respirometry has gained momentum as a rigorous and accurate approach (105). The discovery of new SC assembly factors is also required for the understanding of the molecular pathways leading to SC formation and its plasticity upon changing environmental and nutritional conditions. New structural information, screening for proteins detected in SCs, or Complexome Profiling Alignment (COPAL) to elucidate novel proteins co-migrating with SCs could lead to the detection of new proteins involved in SC assembly and biogenesis. Also, with the structural information available, it is feasible to engineer yeast strains and human cell lines containing fully functional MRC complexes but unable to superassemble, as well as lines in which the MRC complexes are permanently linked. Furthermore, the field requires the development of new methodologies to analyze, in live cells, MRC SCs formation and disruption, to understand whether and how the process changes in a dynamic-manner to respond to nutritional or environmental stressors. Approaches such as fluorescence lifetime imaging microscopy, explained earlier, could be useful to provide answers to some of the remaining questions in the field.

Changes induced in complexes and supercomplexes (SCs) of the mitochondrial respiratory chain under low oxygen levels. CI undergoes a switch from its active to its dormant form, which has been related to destabilization of CI in the SCs in plants. HIF1 induces the expression of hypoxic isoforms of subunits of CIV NDUFA4L2 (SWISS-MODEL Q9NRX3) and COX4-2, while upregulating LON protease that degrades COX4-1 isoform of CIV. There is also a HIF1-dependent induction of HIGD1A (RCSB PDB 2LOM), to generate a more efficient CIV, and a decrease of the CIV subunit COX5B. Concurrently, miR-210 is activated and downregulates ISCU1/2, decreasing the levels of NDUFS subunits of CI and SDHD of CII. CIV subunit NDUFA4 and CIV assembly factor COX10 are also targets of miR-210, lowering their levels.

CI, complex I (dormant state RCSB PDB 5O31; active state RCSB PDB 5XTD). CII, avian complex II (RCSB PDB 1YQ3). CIII, complex III (RCSB PDB 5XTE). CIV, complex IV (RCSB PDB 5X62). SC, respiratory supercomplex I+III2+IV or respirasome (RCSB PDB 5GPN). HIF, Hypoxia Inducible Factor. HRE, Hypoxia Responsive Element. IMS, Intermembrane Space.

Acknowledgements: Our studies are supported by NIH-R35 grant GM118141 to A.B.

References

1. Maxwell, P.H. et al. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 399, 271-275 (1999). https://doi.org/10.1038/20459
PMid:10353251

 

2. Semenza, G.L. & Wang, G.L. A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation. Mol Cell Biol 12, 5447-5454 (1992). https://doi.org/10.1128/MCB.12.12.5447 PMid:1448077 PMCid:PMC360482

 
 

3. Wang, G.L., Jiang, B.H., Rue, E.A. & Semenza, G.L. Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci U S A 92, 5510-5514 (1995). https://doi.org/10.1073/pnas.92.12.5510 PMid:7539918 PMCid:PMC41725

 
 

4. Bracken, C.P. et al. Cell-specific regulation of hypoxia-inducible factor (HIF)-1alpha and HIF-2alpha stabilization and transactivation in a graded oxygen environment. J Biol Chem 281, 22575-22585 (2006). https://doi.org/10.1074/jbc.M600288200 PMid:16760477

 
 

5. Jiang, B.H., Semenza, G.L., Bauer, C. & Marti, H.H. Hypoxia-inducible factor 1 levels vary exponentially over a physiologically relevant range of O2 tension. Am J Physiol 271, C1172-1180 (1996). https://doi.org/10.1152/ajpcell.1996.271.4.C1172 PMid:8897823

 
 

6. Koh, M.Y. & Powis, G. Passing the baton: the HIF switch. Trends Biochem Sci 37, 364-372 (2012). https://doi.org/10.1016/j.tibs.2012.06.004 PMid:22818162 PMCid:PMC3433036

 
 

7. Benita, Y. et al. An integrative genomics approach identifies Hypoxia Inducible Factor-1 (HIF-1)-target genes that form the core response to hypoxia. Nucleic Acids Res 37, 4587-4602 (2009).
https://doi.org/10.1093/nar/gkp425 PMid:19491311 PMCid:PMC2724271

 
 

8. Kim, J.W., Tchernyshyov, I., Semenza, G.L. & Dang, C.V. HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab 3, 177-185 (2006). https://doi.org/10.1016/j.cmet.2006.02.002 PMid:16517405

 
 

9. Chi, J.T. et al. Gene expression programs in response to hypoxia: cell type specificity and prognostic significance in human cancers. PLoS Med 3, e47 (2006). https://doi.org/10.1371/journal.pmed.0030047 PMid:16417408 PMCid:PMC1334226

 
 

10. Xia, X. et al. Integrative analysis of HIF binding and transactivation reveals its role in maintaining histone methylation homeostasis. Proc Natl Acad Sci U S A 106, 4260-4265 (2009).
https://doi.org/10.1073/pnas.0810067106 PMid:19255431 PMCid:PMC2657396

 
 

11. Luo, W., Chang, R., Zhong, J., Pandey, A. & Semenza, G.L. Histone demethylase JMJD2C is a coactivator for hypoxia-inducible factor 1 that is required for breast cancer progression. Proc Natl Acad Sci U S A 109, E3367-3376 (2012). https://doi.org/10.1073/pnas.1217394109 PMid:23129632 PMCid:PMC3523832

 
 

12. Cavadas, M.A.S., Cheong, A. & Taylor, C.T. The regulation of transcriptional repression in hypoxia. Exp Cell Res 356, 173-181 (2017). https://doi.org/10.1016/j.yexcr.2017.02.024
PMid:28219680

 
 

13. Batie, M., Del Peso, L. & Rocha, S. Hypoxia and Chromatin: A Focus on Transcriptional Repression Mechanisms. Biomedicines 6 (2018). https://doi.org/10.20944/preprints201803.0004.v1

 
 

14. Serocki, M. et al. miRNAs regulate the HIF switch during hypoxia: a novel therapeutic target. Angiogenesis 21, 183-202 (2018). https://doi.org/10.1007/s10456-018-9600-2 PMid:29383635 PMCid:PMC5878208

 
 

15. Papandreou, I., Cairns, R.A., Fontana, L., Lim, A.L. & Denko, N.C. HIF-1 mediates adaptation to hypoxia by actively downregulating mitochondrial oxygen consumption. Cell Metab 3, 187-197 (2006). https://doi.org/10.1016/j.cmet.2006.01.012 PMid:16517406

 
 

16. Aragones, J. et al. Deficiency or inhibition of oxygen sensor Phd1 induces hypoxia tolerance by reprogramming basal metabolism. Nat Genet 40, 170-180 (2008). https://doi.org/10.1038/ng.2007.62 PMid:18176562

 
 

17. Ratcliffe, P.J. Oxygen sensing and hypoxia signalling pathways in animals: the implications of physiology for cancer. J Physiol 591, 2027-2042 (2013). https://doi.org/10.1113/jphysiol.2013.251470 PMid:23401619 PMCid:PMC3634517

 
 

18. Fuhrmann, D.C. & Brune, B. Mitochondrial composition and function under the control of hypoxia. Redox Biol 12, 208-215 (2017). https://doi.org/10.1016/j.redox.2017.02.012 PMid:28259101 PMCid:PMC5333533

 
 

19. Thomas, L.W. & Ashcroft, M. Exploring the molecular interface between hypoxia-inducible factor signalling and mitochondria. Cell Mol Life Sci 76, 1759-1777 (2019).
https://doi.org/10.1007/s00018-019-03039-y PMid:30767037 PMCid:PMC6453877

 
 

20. LaGory, E.L. et al. Suppression of PGC-1alpha Is Critical for Reprogramming Oxidative Metabolism in Renal Cell Carcinoma. Cell Rep 12, 116-127 (2015).
https://doi.org/10.1016/j.celrep.2015.07.027 https://doi.org/10.1016/j.celrep.2015.06.006
PMid:26119730 PMCid:PMC4518559

 
 

21. Zhang, H. et al. HIF-1 inhibits mitochondrial biogenesis and cellular respiration in VHL-deficient renal cell carcinoma by repression of C-MYC activity. Cancer Cell 11, 407-420 (2007).
https://doi.org/10.1016/j.ccr.2007.04.001 PMid:17482131

 
 

22. Zhang, H. et al. Mitochondrial autophagy is an HIF-1-dependent adaptive metabolic response to hypoxia. J Biol Chem 283, 10892-10903 (2008). https://doi.org/10.1074/jbc.M800102200
PMid:18281291 PMCid:PMC2447655

 
 

23. Al-Mehdi, A.B. et al. Perinuclear mitochondrial clustering creates an oxidant-rich nuclear domain required for hypoxia-induced transcription. Sci Signal 5, ra47 (2012).
https://doi.org/10.1126/scisignal.2002712 PMid:22763339 PMCid:PMC3565837

 
 

24. Kim, H. et al. Fine-tuning of Drp1/Fis1 availability by AKAP121/Siah2 regulates mitochondrial adaptation to hypoxia. Mol Cell 44, 532-544 (2011). https://doi.org/10.1016/j.molcel.2011.08.045
PMid:22099302 PMCid:PMC3360955

 
 

25. Chiche, J. et al. Hypoxic enlarged mitochondria protect cancer cells from apoptotic stimuli. J Cell Physiol 222, 648-657 (2010). https://doi.org/10.1002/jcp.21984 PMid:19957303

 
 

26. Degn, H. & Wohlrab, H. Measurement of steady-state values of respiration rate and oxidation levels of respiratory pigments at low oxygen tensions. A new technique. Biochim Biophys Acta 245, 347-355 (1971). https://doi.org/10.1016/0005-2728(71)90153-8

 
 

27. Mazat, J.P., Devin, A. & Ransac, S. Modelling mitochondrial ROS production by the respiratory chain. Cell Mol Life Sci (2019). https://doi.org/10.1007/s00018-019-03381-1 PMid:31748915

 
 

28. Murphy, M.P. How mitochondria produce reactive oxygen species. Biochem J 417, 1-13 (2009).
https://doi.org/10.1042/BJ20081386 PMid:19061483 PMCid:PMC2605959

 
 

29. Angelova, P.R. & Abramov, A.Y. Role of mitochondrial ROS in the brain: from physiology to neurodegeneration. FEBS Lett 592, 692-702 (2018). https://doi.org/10.1002/1873-3468.12964
PMid:29292494

 
 

30. Chistiakov, D.A., Shkurat, T.P., Melnichenko, A.A., Grechko, A.V. & Orekhov, A.N. The role of mitochondrial dysfunction in cardiovascular disease: a brief review. Ann Med 50, 121-127 (2018).
https://doi.org/10.1080/07853890.2017.1417631 PMid:29237304

 
 

31. Kumar, N., Qian, W. & Van Houten, B. Sick mitochondria cause telomere damage: implications for disease. Mol Cell Oncol 7, 1678362 (2020). https://doi.org/10.1080/23723556.2019.1678362
PMid:31993494

 
 

32. Liu, X. & Chen, Z. The pathophysiological role of mitochondrial oxidative stress in lung diseases. J Transl Med 15, 207 (2017). https://doi.org/10.1186/s12967-017-1306-5
PMid:29029603 PMCid:PMC5640915

 
 

33. Monsalve, M., Borniquel, S., Valle, I. & Lamas, S. Mitochondrial dysfunction in human pathologies. Front Biosci 12, 1131-1153 (2007). https://doi.org/10.2741/2132
PMid:17127367

 
 

34. Smith, K.A., Waypa, G.B. & Schumacker, P.T. Redox signaling during hypoxia in mammalian cells. Redox Biol 13, 228-234 (2017). https://doi.org/10.1016/j.redox.2017.05.020 PMid:28595160 PMCid:PMC5460738

 
 

35. Cadenas, S. ROS and redox signaling in myocardial ischemia-reperfusion injury and cardioprotection. Free Radic Biol Med 117, 76-89 (2018). https://doi.org/10.1016/j.freeradbiomed.2018.01.024 PMid:29373843

 
 

36. Diebold, L. & Chandel, N.S. Mitochondrial ROS regulation of proliferating cells. Free Radic Biol Med 100, 86-93 (2016). https://doi.org/10.1016/j.freeradbiomed.2016.04.198 PMid:27154978

 
 

37. Fessler, E. et al. A pathway coordinated by DELE1 relays mitochondrial stress to the cytosol. Nature 579, 433-437 (2020). https://doi.org/10.1038/s41586-020-2076-4 PMid:32132706

 
 

38. Guo, X. et al. Mitochondrial stress is relayed to the cytosol by an OMA1-DELE1-HRI pathway. Nature 579, 427-432 (2020). https://doi.org/10.1038/s41586-020-2078-2 PMid:32132707

 
 

39. Lapuente-Brun, E. et al. Supercomplex assembly determines electron flux in the mitochondrial electron transport chain. Science 340, 1567-1570 (2013). https://doi.org/10.1126/science.1230381
PMid:23812712

 
 

40. Schagger, H. & Pfeiffer, K. Supercomplexes in the respiratory chains of yeast and mammalian mitochondria. EMBO J 19, 1777-1783 (2000). https://doi.org/10.1093/emboj/19.8.1777
PMid:10775262 PMCid:PMC302020

 
 

41. Acin-Perez, R., Fernandez-Silva, P., Peleato, M.L., Perez-Martos, A. & Enriquez, J.A. Respiratory active mitochondrial supercomplexes. Mol Cell 32, 529-539 (2008).
https://doi.org/10.1016/j.molcel.2008.10.021 PMid:19026783

 
 

42. Letts, J.A., Fiedorczuk, K., Degliesposti, G., Skehel, M. & Sazanov, L.A. Structures of Respiratory Supercomplex I+III2 Reveal Functional and Conformational Crosstalk. Mol Cell 75, 1131-1146 e1136 (2019). https://doi.org/10.1016/j.molcel.2019.07.022 PMid:31492636 PMCid:PMC6926478

 
 

43. Sousa, J.S., Mills, D.J., Vonck, J. & Kuhlbrandt, W. Functional asymmetry and electron flow in the bovine respirasome. Elife 5 (2016). https://doi.org/10.7554/eLife.21290 PMid:27830641 PMCid:PMC5117854

 
 

44. Gu, J. et al. The architecture of the mammalian respirasome. Nature 537, 639-643 (2016).
https://doi.org/10.1038/nature19359 PMid:27654917

 
 

45. Letts, J.A., Fiedorczuk, K. & Sazanov, L.A. The architecture of respiratory supercomplexes. Nature 537, 644-648 (2016). https://doi.org/10.1038/nature19774 PMid:27654913

 
 

46. Wu, M., Gu, J., Guo, R., Huang, Y. & Yang, M. Structure of Mammalian Respiratory Supercomplex I1III2IV1. Cell 167, 1598-1609 e1510 (2016). https://doi.org/10.1016/j.cell.2016.11.012 PMid:27912063

 
 

47. Guo, R., Zong, S., Wu, M., Gu, J. & Yang, M. Architecture of Human Mitochondrial Respiratory Megacomplex I2III2IV2. Cell 170, 1247-1257 e1212 (2017). https://doi.org/10.1016/j.cell.2017.07.050 PMid:28844695

 
 

48. D'Aurelio, M., Gajewski, C.D., Lenaz, G. & Manfredi, G. Respiratory chain supercomplexes set the threshold for respiration defects in human mtDNA mutant cybrids. Hum Mol Genet 15, 2157-2169 (2006). https://doi.org/10.1093/hmg/ddl141 PMid:16740593

 
 

49. Lopez-Fabuel, I. et al. Mitochondrial respiratory chain disorganization in Parkinson's disease-relevant PINK1 and DJ1 mutants. Neurochem Int 109, 101-105 (2017).
https://doi.org/10.1016/j.neuint.2017.03.023 PMid:28408307

 
 

50. Sun, D., Li, B., Qiu, R., Fang, H. & Lyu, J. Cell Type-Specific Modulation of Respiratory Chain Supercomplex Organization. Int J Mol Sci 17 (2016). https://doi.org/10.3390/ijms17060926
PMid:27338358 PMCid:PMC4926459

 
 

51. Lobo-Jarne, T. & Ugalde, C. Respiratory chain supercomplexes: Structures, function and biogenesis. Semin Cell Dev Biol 76, 179-190 (2018). https://doi.org/10.1016/j.semcdb.2017.07.021
PMid:28743641 PMCid:PMC5780262

 
 

52. Milenkovic, D., Blaza, J.N., Larsson, N.G. & Hirst, J. The Enigma of the Respiratory Chain Supercomplex. Cell Metab 25, 765-776 (2017). https://doi.org/10.1016/j.cmet.2017.03.009
PMid:28380371

 
 

53. Hirst, J. Open questions: respiratory chain supercomplexes-why are they there and what do they do? BMC Biol 16, 111 (2018). https://doi.org/10.1186/s12915-018-0577-5 PMid:30382836 PMCid:PMC6211484

 
 

54. Fedor, J.G. & Hirst, J. Mitochondrial Supercomplexes Do Not Enhance Catalysis by Quinone Channeling. Cell Metab 28, 525-531 e524 (2018). https://doi.org/10.1016/j.cmet.2018.05.024
PMid:29937372 PMCid:PMC6125145

 
 

55. Stuchebrukhov, A., Schafer, J., Berg, J. & Brzezinski, P. Kinetic advantage of forming respiratory supercomplexes. Biochim Biophys Acta Bioenerg 1861, 148193 (2020).
https://doi.org/10.1016/j.bbabio.2020.148193 PMid:32201307

 
 

56. Calvaruso, M.A. et al. Mitochondrial complex III stabilizes complex I in the absence of NDUFS4 to provide partial activity. Hum Mol Genet 21, 115-120 (2012). https://doi.org/10.1093/hmg/ddr446
PMid:21965299

 
 

57. 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, e102817 (2020).
https://doi.org/10.15252/embj.2019102817 PMid:31912925 PMCid:PMC6996572

 
 

58. Maranzana, E., Barbero, G., Falasca, A.I., Lenaz, G. & Genova, M.L. Mitochondrial respiratory supercomplex association limits production of reactive oxygen species from complex I. Antioxid Redox Signal 19, 1469-1480 (2013). https://doi.org/10.1089/ars.2012.4845 PMid:23581604 PMCid:PMC3797460

 
 

59. Gonzalvez, F. et al. Barth syndrome: cellular compensation of mitochondrial dysfunction and apoptosis inhibition due to changes in cardiolipin remodeling linked to tafazzin (TAZ) gene mutation. Biochim Biophys Acta 1832, 1194-1206 (2013). https://doi.org/10.1016/j.bbadis.2013.03.005 PMid:23523468

 
 

60. McKenzie, M., Lazarou, M., Thorburn, D.R. & Ryan, M.T. Mitochondrial respiratory chain supercomplexes are destabilized in Barth Syndrome patients. J Mol Biol 361, 462-469 (2006).
https://doi.org/10.1016/j.jmb.2006.06.057 PMid:16857210

 
 

61. Diaz, F., Enriquez, J.A. & Moraes, C.T. Cells lacking Rieske iron-sulfur protein have a reactive oxygen species-associated decrease in respiratory complexes I and IV. Mol Cell Biol 32, 415-429 (2012). https://doi.org/10.1128/MCB.06051-11 PMid:22106410 PMCid:PMC3255782

 
 

62. Baracca, A. et al. Mitochondrial Complex I decrease is responsible for bioenergetic dysfunction in K-ras transformed cells. Biochim Biophys Acta 1797, 314-323 (2010). https://doi.org/10.1016/j.bbabio.2009.11.006 PMid:19931505

 
 

63. Lenaz, G. et al. Mitochondrial respiratory chain super-complex I-III in physiology and pathology. Biochim Biophys Acta 1797, 633-640 (2010). https://doi.org/10.1016/j.bbabio.2010.01.025
PMid:20116362

 
 

64. Hernansanz-Agustin, P. et al. Acute hypoxia produces a superoxide burst in cells. Free Radic Biol Med 71, 146-156 (2014). https://doi.org/10.1016/j.freeradbiomed.2014.03.011 PMid:24637263

 
 

65. Hernansanz-Agustin, P. et al. Mitochondrial complex I deactivation is related to superoxide production in acute hypoxia. Redox Biol 12, 1040-1051 (2017).
https://doi.org/10.1016/j.redox.2017.04.025 PMid:28511347 PMCid:PMC5430576

 
 

66. Guzy, R.D. et al. Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metab 1, 401-408 (2005). https://doi.org/10.1016/j.cmet.2005.05.001
PMid:16054089

 
 

67. Paddenberg, R. et al. Essential role of complex II of the respiratory chain in hypoxia-induced ROS generation in the pulmonary vasculature. Am J Physiol Lung Cell Mol Physiol 284, L710-719 (2003). https://doi.org/10.1152/ajplung.00149.2002 PMid:12676762

 
 

68. Babot, M., Birch, A., Labarbuta, P. & Galkin, A. Characterisation of the active/de-active transition of mitochondrial complex I. Biochim Biophys Acta 1837, 1083-1092 (2014).
https://doi.org/10.1016/j.bbabio.2014.02.018 PMid:24569053 PMCid:PMC4331042

 
 

69. Drose, S., Stepanova, A. & Galkin, A. Ischemic A/D transition of mitochondrial complex I and its role in ROS generation. Biochim Biophys Acta 1857, 946-957 (2016).
https://doi.org/10.1016/j.bbabio.2015.12.013 PMid:26777588 PMCid:PMC4893024

 
 

70. Roberts, P.G. & Hirst, J. The deactive form of respiratory complex I from mammalian mitochondria is a Na+/H+ antiporter. J Biol Chem 287, 34743-34751 (2012).
https://doi.org/10.1074/jbc.M112.384560 PMid:22854968 PMCid:PMC3464577

 
 

71. Fuhrmann, D.C. et al. Degradation of the mitochondrial complex I assembly factor TMEM126B under chronic hypoxia. Cell Mol Life Sci 75, 3051-3067 (2018). https://doi.org/10.1007/s00018-018-2779-y PMid:29464284

 
 

72. McCormick, R.I. et al. miR-210 is a target of hypoxia-inducible factors 1 and 2 in renal cancer, regulates ISCU and correlates with good prognosis. Br J Cancer 108, 1133-1142 (2013).
https://doi.org/10.1038/bjc.2013.56 PMid:23449350 PMCid:PMC3619073

 
 

73. Ullmann, P. et al. Hypoxia-responsive miR-210 promotes self-renewal capacity of colon tumor-initiating cells by repressing ISCU and by inducing lactate production. Oncotarget 7, 65454-65470 (2016). https://doi.org/10.18632/oncotarget.11772 PMid:27589845 PMCid:PMC5323168

 
 

74. Chan, S.Y. et al. MicroRNA-210 controls mitochondrial metabolism during hypoxia by repressing the iron-sulfur cluster assembly proteins ISCU1/2. Cell Metab 10, 273-284 (2009).
https://doi.org/10.1016/j.cmet.2009.08.015 PMid:19808020 PMCid:PMC2759401

 
 

75. Merlo, A. et al. Identification of a signaling axis HIF-1alpha/microRNA-210/ISCU independent of SDH mutation that defines a subgroup of head and neck paragangliomas. J Clin Endocrinol Metab 97, E2194-2200 (2012). https://doi.org/10.1210/jc.2012-2410 PMid:22977270

 
 

76. Acin-Perez, R., Gatti, D.L., 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). https://doi.org/10.1016/j.cmet.2011.03.024 PMid:21641552 PMCid:PMC3118639

 
 

77. Arnold, S. & Kadenbach, B. Cell respiration is controlled by ATP, an allosteric inhibitor of cytochrome-c oxidase. Eur J Biochem 249, 350-354 (1997). https://doi.org/10.1111/j.1432-1033.1997.t01-1-00350.x PMid:9363790

 
 

78. Hess, K.C. et al. A mitochondrial CO2-adenylyl cyclase-cAMP signalosome controls yeast normoxic cytochrome c oxidase activity. FASEB J 28, 4369-4380 (2014). https://doi.org/10.1096/fj.14-252890 PMid:25002117 PMCid:PMC4202101

 
 

79. Burke, P.V., Raitt, D.C., Allen, L.A., Kellogg, E.A. & Poyton, R.O. Effects of oxygen concentration on the expression of cytochrome c and cytochrome c oxidase genes in yeast. J Biol Chem 272, 14705-14712 (1997). https://doi.org/10.1074/jbc.272.23.14705 PMid:9169434

 
 

80. Poyton, R.O. Models for oxygen sensing in yeast: implications for oxygen-regulated gene expression in higher eucaryotes. Respir Physiol 115, 119-133 (1999).
https://doi.org/10.1016/S0034-5687(99)00028-6

 
 

81. Huttemann, M., Kadenbach, B. & Grossman, L.I. Mammalian subunit IV isoforms of cytochrome c oxidase. Gene 267, 111-123 (2001). https://doi.org/10.1016/S0378-1119(01)00385-7

 
 

82. Allen, L.A., Zhao, X.J., Caughey, W. & Poyton, R.O. Isoforms of yeast cytochrome c oxidase subunit V affect the binuclear reaction center and alter the kinetics of interaction with the isoforms of yeast cytochrome c. J Biol Chem 270, 110-118 (1995). https://doi.org/10.1074/jbc.270.1.110
PMid:7814361

 
 

83. Fukuda, R. et al. HIF-1 regulates cytochrome oxidase subunits to optimize efficiency of respiration in hypoxic cells. Cell 129, 111-122 (2007). https://doi.org/10.1016/j.cell.2007.01.047
PMid:17418790

 
 

84. Huttemann, M. et al. Cytochrome c oxidase subunit 4 isoform 2-knockout mice show reduced enzyme activity, airway hyporeactivity, and lung pathology. FASEB J 26, 3916-3930 (2012).
https://doi.org/10.1096/fj.11-203273 PMid:22730437 PMCid:PMC3425824

 
 

85. Huttemann, M., Lee, I., Liu, J. & Grossman, L.I. Transcription of mammalian cytochrome c oxidase subunit IV-2 is controlled by a novel conserved oxygen responsive element. FEBS J 274, 5737-5748 (2007). https://doi.org/10.1111/j.1742-4658.2007.06093.x PMid:17937768

 
 

86. Hayashi, T. et al. Higd1a is a positive regulator of cytochrome c oxidase. Proc Natl Acad Sci U S A 112, 1553-1558 (2015). https://doi.org/10.1073/pnas.1419767112 PMid:25605899 PMCid:PMC4321285

 
 

87. Puissegur, M.P. et al. miR-210 is overexpressed in late stages of lung cancer and mediates mitochondrial alterations associated with modulation of HIF-1 activity. Cell Death Differ 18, 465-478 (2011). https://doi.org/10.1038/cdd.2010.119 PMid:20885442 PMCid:PMC3131992

 
 

88. Tello, D. et al. Induction of the mitochondrial NDUFA4L2 protein by HIF-1alpha decreases oxygen consumption by inhibiting Complex I activity. Cell Metab 14, 768-779 (2011).
https://doi.org/10.1016/j.cmet.2011.10.008 PMid:22100406

 
 

89. Lai, R.K. et al. NDUFA4L2 Fine-tunes Oxidative Stress in Hepatocellular Carcinoma. Clin Cancer Res 22, 3105-3117 (2016). https://doi.org/10.1158/1078-0432.CCR-15-1987
PMid:26819450

 
 

90. Soro-Arnaiz, I. et al. Role of Mitochondrial Complex IV in Age-Dependent Obesity. Cell Rep 16, 2991-3002 (2016). https://doi.org/10.1016/j.celrep.2016.08.041 PMid:27626667

 
 

91. Acin-Perez, R. & Enriquez, J.A. The function of the respiratory supercomplexes: the plasticity model. Biochim Biophys Acta 1837, 444-450 (2014). https://doi.org/10.1016/j.bbabio.2013.12.009
PMid:24368156

 
 

92. Greggio, C. et al. Enhanced Respiratory Chain Supercomplex Formation in Response to Exercise in Human Skeletal Muscle. Cell Metab 25, 301-311 (2017).
https://doi.org/10.1016/j.cmet.2016.11.004 PMid:27916530

 
 

93. Huertas, J.R., Al Fazazi, S., Hidalgo-Gutierrez, A., Lopez, L.C. & Casuso, R.A. Antioxidant effect of exercise: Exploring the role of the mitochondrial complex I superassembly. Redox Biol 13, 477-481 (2017). https://doi.org/10.1016/j.redox.2017.07.009 PMid:28719865 PMCid:PMC5512182

 
 

94. Balsa, E. et al. ER and Nutrient Stress Promote Assembly of Respiratory Chain Supercomplexes through the PERK-eIF2alpha Axis. Mol Cell 74, 877-890 e876 (2019).
https://doi.org/10.1016/j.molcel.2019.03.031 PMid:31023583

 
 

95. Gomez, L.A., Monette, J.S., Chavez, J.D., Maier, C.S. & Hagen, T.M. Supercomplexes of the mitochondrial electron transport chain decline in the aging rat heart. Arch Biochem Biophys 490, 30-35 (2009). https://doi.org/10.1016/j.abb.2009.08.002 PMid:19679098 PMCid:PMC2762268

 
 

96. Ramirez-Camacho, I., Flores-Herrera, O. & Zazueta, C. The relevance of the supramolecular arrangements of the respiratory chain complexes in human diseases and aging. Mitochondrion 47, 266-272 (2019). https://doi.org/10.1016/j.mito.2019.01.001 PMid:30664953

 
 

97. Jang, S. & Javadov, S. Current Challenges in Elucidating Respiratory Supercomplexes in Mitochondria: Methodological Obstacles. Front Physiol 9, 238 (2018). https://doi.org/10.3389/fphys.2018.00238 PMid:29615931 PMCid:PMC5864997

 
 

98. Varnaite, R. & MacNeill, S.A. Meet the neighbors: Mapping local protein interactomes by proximity-dependent labeling with BioID. Proteomics 16, 2503-2518 (2016).
https://doi.org/10.1002/pmic.201670171 https://doi.org/10.1002/pmic.201600123 PMid:27329485 PMCid:PMC5053326

 
 

99. Chavez, J.D. et al. Chemical Crosslinking Mass Spectrometry Analysis of Protein Conformations and Supercomplexes in Heart Tissue. Cell Syst 6, 136-141 e135 (2018).
https://doi.org/10.1016/j.cels.2017.10.017 PMid:29199018 PMCid:PMC5799023

 
 

100. Liu, F., Lossl, P., Rabbitts, B.M., Balaban, R.S. & Heck, A.J.R. The interactome of intact mitochondria by cross-linking mass spectrometry provides evidence for coexisting respiratory supercomplexes. Mol Cell Proteomics 17, 216-232 (2018). https://doi.org/10.1074/mcp.RA117.000470 PMid:29222160 PMCid:PMC5795388

 
 

101. Schweppe, D.K. et al. Mitochondrial protein interactome elucidated by chemical cross-linking mass spectrometry. Proc Natl Acad Sci U S A 114, 1732-1737 (2017).
https://doi.org/10.1073/pnas.1617220114 PMid:28130547 PMCid:PMC5321032

 
 

102. Rieger, B. et al. Lifetime imaging of GFP at CoxVIIIa reports respiratory supercomplex assembly in live cells. Sci Rep 7, 46055 (2017). https://doi.org/10.1038/srep46055 PMid:28383048 PMCid:PMC5382582

 
 

103. Ramirez-Aguilar, S.J. et al. The composition of plant mitochondrial supercomplexes changes with oxygen availability. J Biol Chem 286, 43045-43053 (2011). https://doi.org/10.1074/jbc.M111.252544 PMid:22009743 PMCid:PMC3234806

 
 

104. Bundgaard, A., James, A.M., Joyce, W., Murphy, M.P. & Fago, A. Suppression of reactive oxygen species generation in heart mitochondria from anoxic turtles: the role of complex I S-nitrosation. J Exp Biol 221 (2018). https://doi.org/10.1242/jeb.174391 PMid:29496783 PMCid:PMC5963835

 
 

105. 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 e1784 (2018). https://doi.org/10.1016/j.celrep.2018.10.058 PMid:30428348 PMCid:PMC6286155

 
 

106. Perez-Perez, 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). https://doi.org/10.1016/j.celrep.2016.07.081 PMid:27545886 PMCid:PMC5007171

 
 

107. Cogliati, S. et al. Mechanism of super-assembly of respiratory complexes III and IV. Nature 539, 579-582 (2016). https://doi.org/10.1038/nature20157 PMid:27775717

 
 

108. Williams, E.G. et al. Systems proteomics of liver mitochondria function. Science 352, aad0189 (2016). https://doi.org/10.1126/science.aad0189 PMid:27284200

 
 

109. Watanabe, T. et al. Isolation of estrogen-responsive genes with a CpG island library. Mol Cell Biol 18, 442-449 (1998). https://doi.org/10.1128/MCB.18.1.442 PMid:9418891 PMCid:PMC121513

 
 

110. Ikeda, K., Shiba, S., Horie-Inoue, K., Shimokata, K. & Inoue, S. A stabilizing factor for mitochondrial respiratory supercomplex assembly regulates energy metabolism in muscle. Nat Commun 4, 2147 (2013). https://doi.org/10.1038/ncomms3147 PMid:23857330

 
 

111. Rossignol, R. et al. Energy substrate modulates mitochondrial structure and oxidative capacity in cancer cells. Cancer Res 64, 985-993 (2004). https://doi.org/10.1158/0008-5472.CAN-03-1101
PMid:14871829

 
 

112. Shiba, S. et al. Deficiency of COX7RP, a mitochondrial supercomplex assembly promoting factor, lowers blood glucose level in mice. Sci Rep 7, 7606 (2017). https://doi.org/10.1038/s41598-017-08081-z PMid:28790391 PMCid:PMC5548899

 
 

113. Ikeda, K. et al. Mitochondrial supercomplex assembly promotes breast and endometrial tumorigenesis by metabolic alterations and enhanced hypoxia tolerance. Nat Commun 10, 4108 (2019). https://doi.org/10.1038/s41467-019-12124-6 PMid:31511525 PMCid:PMC6739376

 
 

114. Chen, Y.C. et al. Identification of a protein mediating respiratory supercomplex stability. Cell Metab 15, 348-360 (2012). https://doi.org/10.1016/j.cmet.2012.02.006 PMid:22405070 PMCid:PMC3302151

 
 

115. Vukotic, M. et al. Rcf1 mediates cytochrome oxidase assembly and respirasome formation, revealing heterogeneity of the enzyme complex. Cell Metab 15, 336-347 (2012).
https://doi.org/10.1016/j.cmet.2012.01.016 PMid:22342701

 
 

116. Ameri, K. et al. HIGD1A Regulates Oxygen Consumption, ROS Production, and AMPK Activity during Glucose Deprivation to Modulate Cell Survival and Tumor Growth. Cell Rep (2015).
https://doi.org/10.1016/j.celrep.2015.01.020 PMid:25683712 PMCid:PMC4534363

 
 

117. An, H.J. et al. The survival effect of mitochondrial Higd-1a is associated with suppression of cytochrome C release and prevention of caspase activation. Biochim Biophys Acta 1813, 2088-2098 (2011). https://doi.org/10.1016/j.bbamcr.2011.07.017 PMid:21856340

 
 

118. Vidoni, S. et al. MR-1S Interacts with PET100 and PET117 in Module-Based Assembly of Human Cytochrome c Oxidase. Cell Rep 18, 1727-1738 (2017).
https://doi.org/10.1016/j.celrep.2017.01.044 PMid:28199844

 
 

119. Timon-Gomez, A., Garlich, J., Stuart, R.A., Ugalde, C. & Barrientos, A. Distinct Roles of Mitochondrial HIGD1A and HIGD2A in Respiratory Complex and Supercomplex Biogenesis. Cell Rep 31, 107607 (2020). https://doi.org/10.1016/j.celrep.2020.107607 PMid:32375044