Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 An overview of cytochrome c oxidase assembly factors and some hypotheses about the assembly process + 2617 word(s) 2617 2020-10-07 05:08:53 |
2 format correct Meta information modification 2617 2020-10-10 11:31:16 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Watson, S.A.; Mcstay, G.P. Cytochrome c oxidase Assembly Factors. Encyclopedia. Available online: https://encyclopedia.pub/entry/2387 (accessed on 20 April 2024).
Watson SA, Mcstay GP. Cytochrome c oxidase Assembly Factors. Encyclopedia. Available at: https://encyclopedia.pub/entry/2387. Accessed April 20, 2024.
Watson, Shane A., Gavin P. Mcstay. "Cytochrome c oxidase Assembly Factors" Encyclopedia, https://encyclopedia.pub/entry/2387 (accessed April 20, 2024).
Watson, S.A., & Mcstay, G.P. (2020, October 07). Cytochrome c oxidase Assembly Factors. In Encyclopedia. https://encyclopedia.pub/entry/2387
Watson, Shane A. and Gavin P. Mcstay. "Cytochrome c oxidase Assembly Factors." Encyclopedia. Web. 07 October, 2020.
Cytochrome c oxidase Assembly Factors
Edit

An overview of cytochrome c oxidase assembly factors in baker's yeast and humans with hypotheses regarding the assembly process in low oxygen conditions and during initial entry into the inner mitochondrial membrane.

mitochondria cytochrome c oxidase oxidative phosphorylation electron transport chain

1. Introduction

Energy generation is a fundamental process that supports all forms of life on earth. The universal energy currency in life is adenosine triphosphate (ATP) which captures energy from bioenergetic and metabolic processes powered by substrate-level and oxidative phosphorylation. In eukaryotes a major portion of ATP is generated through mitochondria, a double-membrane organelle derived from a symbiotic relationship at the origin of multicellular life [1]. Mitochondria harbour a series of multi-subunit complexes that perform electron transfer and proton translocation from the internal mitochondrial matrix space to the inter-membrane space (IMS) through the inner mitochondrial (IMM). This generates an electrochemical gradient across the IMM where proton concentration is higher in the IMS than the matrix. The accumulated protons can re-enter the matrix through ATP synthase - a proton transporter that couples the proton gradient to the synthesis of ATP. The electron transport complexes are an ordered series of multi-subunit complexes that accept electrons from carriers such as nicotinamide adenine dinucleotide (NAD+/NADH) and flavin adenine dinucleotide (FAD/FADH2) produced through metabolic pathways such as glycolysis, citric acid cycle and β-oxidation. NADH is oxidised by NADH dehydrogenase enzymes, such as multi-subunit Complex I in many eukaryotes or single subunit Ndi1p in Saccharomyces cerevisiae. The first step in Complex I sees the electrons passing through a series of co-factors and prosthetic groups such as iron-sulphur (Fe-S) clusters and flavin mononucleotide (FMN). The electrons are passed to the membrane soluble lipid based redox carrier co-enzyme Q (ubiquinone/ubiquinol). In addition, FAD/FADH2 is reduced by succinate dehydrogenase of the citric acid cycle which also acts as complex II of the electron transport chain (ETC). Complex II contains Fe-S clusters and haem prosthetic groups to again pass electrons to Co-enzyme Q. Reduced Co-enzyme Q (ubiquinol) is oxidised by complex III (co-enzyme Q:cytochrome c oxidoreductase) using Fe-S and haem groups to reduce cytochrome c, the next mobile electron carrier in the ETC. Cytochrome c localises to the IMS face of the IMM to shuttle electrons from complex III to complex IV using a haem prosthetic group. Complex IV (cytochrome c oxidase) is the final complex of the ETC catalysing the reduction of molecular oxygen, the terminal electron acceptor to form water via copper and haem groups. The process of electron transfer is coupled to proton translocation from the matrix to the IMS and is mediated by complexes I, III and IV to generate the electrochemical gradient.

The ETC complexes are multi-subunit assemblies with protein subunits derived from both nuclear and mitochondrial gene products. To ensure efficient assembly of complexes co-ordinated post-translational processing and protein-protein interactions are required. This is accomplished by a stepwise biogenesis pathway that requires the function of specific chaperones. As oxidative phosphorylation (OXPHOS) is an evolutionarily conserved process in all eukaryotes much understanding of the assembly process can be gleaned from the study from all eukaryotic organisms where mitochondria exist. This is true for single cell eukaryotes such as the budding yeast Saccharomyces cerevisiae and photosynthetic alga Chlamydomoas reinhardtii to multicellular higher eukaryotes such as human and mouse. The combination of genetic manipulation of eukaryotic model organisms and naturally occurring mutations in human disease has provided much information we have at the moment about how assembly of the ETC occurs. The molecular details of the most studied mammalian organisms and yeast OXPHOS complex structure and assembly have very many similarities but also some differences. These differences are in the number of structural subunits forming complexes as well as the post-translational processing required for correct assembly. Furthermore, multicellular eukaryotes express tissue-specific subunits that provide unique functions to the complexes. Budding yeast has been a powerful model system to understand assembly of OXPHOS complexes as they are facultative anaerobes, meaning they can sustain energy requirements in the absence of functional OXPHOS if grown on a carbon source, such as glucose, that does not require the functions of OXPHOS. When grown on carbon sources such as ethanol, glycerol and lactate, that require functional OXPHOS, any defects in assembly or function of the ETC or OXPHOS results in a lack of growth. This has enabled a large number of genes to be identified that are responsible for ETC and OXPHOS assembly and function through powerful genetic and phenotypic analysis [2]. Recently, more homologues have been identified in higher eukaryotes demonstrating the conservation of the assembly of the complex as well as the enzymatic function [3].

2. Cytochrome c oxidase Structure and Function

In order to understand the assembly pathway of multi-subunit complexes it is important to put this in the context of the fully assembled and functional complex. Each subunit has a defined three-dimensional organisation and interacts with certain subunits, both required for full function. Mammalian and yeast cytochrome c oxidase (COX) are composed of 3 subunits derived from mitochondrial DNA and between 8 and 11 subunits derived from nuclear genes. The mitochondrial encoded subunits Cox1p and Cox2p contain prosthetic groups and co-factors required for electron transfer including haem groups in Cox1p and copper ions in Cox1p and Cox2p. These two subunits are also responsible for the translocation of protons from the matrix to the IMS through hydrophilic channels. The post-translational associations of prosthetic groups and co-factors as well as proteolytic processing and membrane insertion are accomplished by a number of chaperones, some specific to COX and others shared with other OXPHOS complexes. All of these chaperones are encoded by the nuclear genome and for the most part functional homologs are conserved from budding yeast to higher mammals with some exceptions. The functions of many of these chaperones have been determined and are conserved; however, there are still some genes without assigned functions or functional homologs [4]. Determination of gene function and functional homologs will allow for a more complete understanding of the process of COX biogenesis.

Defects in OXPHOS are responsible for an array of genetic disorders that impact on tissues and organs with high metabolic demands or dependency on mitochondrial metabolic pathways. Specifically for COX there are a number of mutations in assembly factor and structural genes that cause forms of Leigh Syndrome, mitochondrial Complex IV deficiency and rare syndromes that result in neurological disorders along with impacts on the heart [5], liver [6], kidney [7] and digestive system [8]. The neurological disorders differ in their onset severity depending on the nature of the mutation. For example, encephalopathies display clinical features due to specific mutations in all OXPHOS complexes as well as other mitochondrial genes involved in mitochondrial DNA maintenance, e.g., DNA polymerase γ [9]. Variation in severity is also observed by certain combinations of clinical features, for example, encephalopathies presenting with tubulopathy and hepatopathy, caused by mutations in the Complex III gene BCS1L [10]. Specific mutations associated with these diseases are individually rare, but when combined with other diseases that cause defects in OXPHOS occur at approximately 1 in 5000 births, representing one of the highest incidences of inherited metabolic diseases affecting humans. Through recent advances in diagnosis of mitochondrial disease and investigation of the molecular basis of the disease insights into the functions of assembly factors is improving [11].

In humans and other higher mammals, most assembly factors and structural subunits have an orthologue in budding yeast. The human COX enzyme is composed of 14 subunits, which is 3 more than budding yeast. As with most of the nuclear encoded structural subunits these are single trans-membrane spanning polypeptides that surround the core catalytic subunits and do not interfere with conserved interfaces and interactions [12]. Recently, other subunits have been identified as associated with purified COX including the hypoxia inducible gene Rcf1p/HIGD2A which acts as a link between Complex III to support supercomplex formation especially under anaerobic conditions when Cox5bp is expressed [13][14]. The structure of mammalian COX has been known since the 1980s and recently the 14th subunit was identified. NDUFA4 was originally attributed to Complex I, but biochemical and genetic studies revealed deficiencies in COX activity and an equal stoichiometry with other COX structural subunits without impact on Complex I activity [15][16][17]. NDUFA4 has been identified in purified COX containing supercomplex structures determined by X-ray crystallography at the interface of COX dimers [18]. This is further supported by the presence of a homologue in budding yeast, MIN8/YPR010C-A, identified through proteomics [19]. S. cerevisiae do not express a multi-subunit Complex I NADH dehydrogenase, fully supporting another role for this gene. It is annotated as a mitochondrial gene of unknown function. Further work is required to understand the function of this gene as this has not been identified as a gene required for respiratory deficiency in previous studies. The function of these extra subunits is subject to speculation. S. cerevisiae COX can retain function in the absence of the Cox8p and Cox13p subunits indicating that not all structural subunits are required for activity [20]. COX8A is the human homologue of yeast Cox8p, and it is the smallest subunit of human COX. In contrast to S.cerevisiae, a mutation in this subunit is responsible for neurological disorder due to loss of COX function [21]. S. cerevisiae COX13 has two homologues in humans-COX6A1 and COX6A2-tissue specific isoforms in the liver and striated muscle, respectively. Mutations in the isoform COX6A1 have been shown to be the cause of Charcot-Marie tooth disease, a neuropathological disorder [22][23]. Mutations in the second isoform COX6A2 can lead to muscle specific Complex IV deficiency [24]. These discrepancies indicate differences in the essential nature of these peripheral subunits. The 3 human specific subunits in human COX are COX6C, COX7C and NDUFA4. Mutations in NDUFA4 are associated with Leigh Disease [25], while mutations in the two other subunits COX6C and COX7C have not been associated with disease. Even though there are differences between the phenotypes of yeast and humans, this should be taken with some caution as some homologous subunits do not have the same common origin [26].

3. Cytochrome c oxidase Assembly Factors

Biogenesis of multi-subunit complexes requires specific maturation steps for subunits to ensure correct localisation and topology, post-translational modifications and interactions with other subunits. These biogenesis pathways also require quality control points to verify each step has been correctly performed but also to determine whether downstream steps can occur. Assembly factors for COX can be broadly grouped by function including transcription and translational regulation, membrane insertion, proteolysis, co-factor and prosthetic association or an as yet undefined function. The process of COX assembly ensures appropriate association of correctly matured subunits with incorporated co-factors and prosthetic groups when in adequate abundance. This prevents accumulation of mid-assembly intermediates that can cause potentially detrimental effects due to excess reactive oxygen species (ROS) or overloading of protein quality control systems. Due to the number of subunits that need to associate in the IMM, and the number of maturation steps required, there are many assembly factors to ensure each of these steps are performed correctly. The assembly pathway still needs further elucidation to assign functions to specific assembly factors and how certain assembly intermediates interact, but general functions are beginning to be understood.

4. Mutations in Cytochrome c oxidase Assembly Factors as Cause of Human Disease

Understanding of COX assembly has also been captured from the study of patients suffering from disease caused by mutations in these genes. Mutations in genes could result in the absence of assembled COX but also in mutation in critical residues for the activity of the enzyme. The inheritance pattern of these diseases also contributes to the severity of the disease. This is why mitochondrial diseases present as a heterogeneous spectrum of disorders that are difficult to diagnose and treat. A number of mutations in assembly factors are the cause of these diseases [11]. The most common cause of complex IV deficiency is through mutation of SURF1. There are at least 36 different mutations in this gene that result in disease, these are located throughout the length of the gene and result in missense and nonsense mutations. Some of these have been introduced into budding yeast SHY1 at conserved sites and result in the accumulation of intermediates at different stages of Cox1p module assembly. This indicates that Shy1p function can be sensed by the Mss51p checkpoint, but has functions later in the assembly process [27]. The wide variety of disease causing SURF1 mutations, along with the evidence of a weak and spontaneously suppressible respiratory defect in budding yeast, indicate that defects in SURF1 can be overcome by changes in a variety of other genes in the Cox1 assembly pathway. A similar explanation may be behind the number of mutations identified in the copper chaperones, COX10 and COX15 [28][29][30]. There are a number of homologues in this pathway that if altered in expression could explain the disease. Furthermore, as elevated copper can suppress yeast mutations, alterations in cellular copper handling could explain how a mutation is overcome to enable viability. Mutations in other genes are much less frequent and this is most likely explained by a defect that cannot be overcome by changes in gene expression. This also indicates that during embryonic development there is an essential function lacking when a mutation in other genes are present that cannot be overcome and the embryo is not viable. There are several examples where genes harbouring mutations lead to COX deficiency, but the role has not been fully characterised. CEP89 was identified as the mutated gene in a patient with COX deficiency. This mutation caused loss of COX activity and function [31]. The exact role of CEP89 in this process still has to be determined. In contrast to S. cerevisiae, a point mutation in COX14 results in decreased synthesis of COX1 with an expected loss of COX expression and activity potentially highlighting a divergence in the assembly process [32]. Mutations in COA3, which functions at a similar point to COX14, also result in a mitochondrial disease due to specific loss of COX expression [33]. Mutations in COA5, the human homologue of Pet191p, results in infantile cardioencephalomyopathy, caused by COX deficiency. Analysis of native complexes demonstrated a Cox1 containing assembly intermediate that may represent MITRAC [34]. COA7 mutants result in COX deficiency causing leukoencephalopathies and peripheral neuropathies [35][36] and this deficiency can be restored by inhibiting cytosolic degradation of COA7 indicating these mutations delay COA7 import into mitochondria which are still capable of contributing to COX assembly [37]. Mutations in PET100 also cause COX deficiency through a truncation and an import defect [38][39][40]. PET117 mutations cause COX deficiency, most likely through its role in assisting haem synthesis coupling to association into Cox1 [41]. In patients with COX10 mutations COX expression is decreased and Cox1 exists in a sub-assembly that migrates similar to MITRAC [42][43]. COX15 mutations also show similar deficiencies of COX; however, the assembly intermediates in these patients have not been investigated [44][30][45][46][47]. Mutations in the copper handling proteins, SCO1 and SCO2, have also been identified resulting in COX deficiency [48][49][50][51][52][53]. COA8, a less characterised protein identified as having a role in programmed cell death, when mutated can also cause COX deficiency [54].

5. Summary

In summary, the assembly of COX is a conserved process that requires the essential function of many assembly factors, somewhat unexpectedly more than the number of structural genes. This indicates the complexity and importance of the process. Without a functional COX enzyme, mitochondria are not able to produce ATP which has devastating consequences in the context of severely debilitating diseases that often lead to early death. Through further understanding of how COX assembly occurs better points of diagnostic and therapeutic intervention will be developed to improve the quality of life of patients suffering from these diseases, and families facing choices that cause anxiety and distress.

References

  1. Gray, M.W. Mitochondrial evolution. Cold Spring Harb. Perspect. Biol. 2012, 4, a011403.
  2. Tzagoloff, A.; Dieckmann, C.L. PET genes of Saccharomyces cerevisiae. Microbiol. Rev. 1990, 54, 211–225.
  3. Barros, M.H.; McStay, G.P. Modular biogenesis of mitochondrial respiratory complexes. Mitochondrion 2020, 50, 94–114.
  4. Mick, D.U.; Fox, T.D.; Rehling, P. Inventory control: Cytochrome c oxidase assembly regulates mitochondrial translation. Nat. Rev. Mol. Cell Biol. 2011, 12, 14–20.
  5. Limongelli, G.; Masarone, D.; Pacileo, G. Mitochondrial disease and the heart. Heart 2017, 103, 390–398.
  6. Lee, W.S.; Sokol, R.J. Mitochondrial hepatopathies: Advances in genetics, therapeutic approaches, and outcomes. J. Pediatr. 2013, 163, 942–948.
  7. O’Toole, J.F. Renal manifestations of genetic mitochondrial disease. Int. J. Nephrol. Renovasc. Dis. 2014, 7, 57–67.
  8. Finsterer, J.; Frank, M. Gastrointestinal manifestations of mitochondrial disorders: A systematic review. Ther. Adv. Gastroenterol. 2017, 10, 142–154.
  9. McFarland, R.; Taylor, R.W.; Turnbull, D.M. A neurological perspective on mitochondrial disease. Lancet Neurol. 2010, 9, 829–840.
  10. Visapää, I.; Fellman, V.; Vesa, J.; Dasvarma, A.; Hutton, J.L.; Kumar, V.; Payne, G.S.; Makarow, M.; Van Coster, R.; Taylor, R.W.; et al. GRACILE syndrome, a lethal metabolic disorder with iron overload, is caused by a point mutation in BCS1L. Am. J. Hum. Genet. 2002, 71, 863–876.
  11. Gorman, G.S.; Chinnery, P.F.; DiMauro, S.; Hirano, M.; Koga, Y.; McFarland, R.; Suomalainen, A.; Thorburn, D.R.; Zeviani, M.; Turnbull, D.M. Mitochondrial diseases. Nat. Rev. Dis. Primers 2016, 2, 16080.
  12. Maréchal, A.; Meunier, B.; Lee, D.; Orengo, C.; Rich, P.R. Yeast cytochrome c oxidase: A model system to study mitochondrial forms of the haem-copper oxidase superfamily. Biochim. Biophys. Acta 2012, 1817, 620–628.
  13. Hartley, A.M.; Meunier, B.; Pinotsis, N.; Maréchal, A. Rcf2 revealed in cryo-EM structures of hypoxic isoforms of mature mitochondrial III-IV supercomplexes. Proc. Natl. Acad. Sci. USA 2020, 117, 9329–9337.
  14. Timón-Gómez, 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. 2020, 31, 107607.
  15. Balsa, E.; Marco, R.; Perales-Clemente, E.; Szklarczyk, R.; Calvo, E.; Landázuri, M.O.; Enríquez, J.A. NDUFA4 is a subunit of complex IV of the mammalian electron transport chain. Cell Metab. 2012, 16, 378–386.
  16. Pitceathly, R.D.S.; Taanman, J.-W. NDUFA4 (Renamed COXFA4) Is a Cytochrome-c Oxidase Subunit. Trends Endocrinol. Metab. 2018, 29, 452–454.
  17. Kadenbach, B. Regulation of Mammalian 13-Subunit Cytochrome c Oxidase and Binding of other Proteins: Role of NDUFA4. Trends Endocrinol. Metab. 2017.
  18. Zong, S.; Wu, M.; Gu, J.; Liu, T.; Guo, R.; Yang, M. Structure of the intact 14-subunit human cytochrome c oxidase. Cell Res. 2018, 28, 1026–1034.
  19. Morgenstern, M.; Stiller, S.B.; Lübbert, P.; Peikert, C.D.; Dannenmaier, S.; Drepper, F.; Weill, U.; Höß, P.; Feuerstein, R.; Gebert, M.; et al. Definition of a High-Confidence Mitochondrial Proteome at Quantitative Scale. Cell Rep. 2017, 19, 2836–2852.
  20. McStay, G.P.; Su, C.-H.; Thomas, S.M.; Xu, J.T.; Tzagoloff, A. Characterization of assembly intermediates containing subunit 1 of yeast cytochrome oxidase. J. Biol. Chem. 2013, 288, 26546–26556.
  21. Hallmann, K.; Kudin, A.P.; Zsurka, G.; Kornblum, C.; Reimann, J.; Stüve, B.; Waltz, S.; Hattingen, E.; Thiele, H.; Nürnberg, P.; et al. Loss of the smallest subunit of cytochrome c oxidase, COX8A, causes Leigh-like syndrome and epilepsy. Brain 2016, 139, 338–345.
  22. Tamiya, G.; Makino, S.; Hayashi, M.; Abe, A.; Numakura, C.; Ueki, M.; Tanaka, A.; Ito, C.; Toshimori, K.; Ogawa, N.; et al. A mutation of COX6A1 causes a recessive axonal or mixed form of Charcot-Marie-Tooth disease. Am. J. Hum. Genet. 2014, 95, 294–300.
  23. Laššuthová, P.; Beharka, R.; Krůtová, M.; Neupauerová, J.; Seeman, P. COX6A1 mutation causes axonal hereditary motor and sensory neuropathy—The confirmation of the primary report. Clin. Genet. 2016, 89, 512–514.
  24. Inoue, M.; Uchino, S.; Iida, A.; Noguchi, S.; Hayashi, S.; Takahashi, T.; Fujii, K.; Komaki, H.; Takeshita, E.; Nonaka, I.; et al. COX6A2 variants cause a muscle-specific cytochrome c oxidase deficiency. Ann. Neurol. 2019.
  25. Pitceathly, R.D.S.; Rahman, S.; Wedatilake, Y.; Polke, J.M.; Cirak, S.; Foley, A.R.; Sailer, A.; Hurles, M.E.; Stalker, J.; Hargreaves, I.; et al. UK10K Consortium NDUFA4 mutations underlie dysfunction of a cytochrome c oxidase subunit linked to human neurological disease. Cell Rep. 2013, 3, 1795–1805.
  26. Pierron, D.; Wildman, D.E.; Hüttemann, M.; Markondapatnaikuni, G.C.; Aras, S.; Grossman, L.I. Cytochrome c oxidase: Evolution of control via nuclear subunit addition. Biochim. Biophys. Acta 2012, 1817, 590–597.
  27. Reinhold, R.; Bareth, B.; Balleininger, M.; Wissel, M.; Rehling, P.; Mick, D.U. Mimicking a SURF1 allele reveals uncoupling of cytochrome c oxidase assembly from translational regulation in yeast. Hum. Mol. Genet. 2011, 20, 2379–2393.
  28. Valnot, I.; von Kleist-Retzow, J.C.; Barrientos, A.; Gorbatyuk, M.; Taanman, J.W.; Mehaye, B.; Rustin, P.; Tzagoloff, A.; Munnich, A.; Rötig, A. A mutation in the human heme A: Farnesyltransferase gene (COX10 ) causes cytochrome c oxidase deficiency. Hum. Mol. Genet. 2000, 9, 1245–1249.
  29. Sinkler, C.A.; Kalpage, H.; Shay, J.; Lee, I.; Malek, M.H.; Grossman, L.I.; Hüttemann, M. Tissue- and Condition-Specific Isoforms of Mammalian Cytochrome c Oxidase Subunits: From Function to Human Disease. Oxid. Med. Cell. Longev. 2017, 2017, 1534056.
  30. Halperin, D.; Drabkin, M.; Wormser, O.; Yogev, Y.; Dolgin, V.; Shorer, Z.; Gradstein, L.; Shelef, I.; Flusser, H.; Birk, O.S. Phenotypic variability and mutation hotspot in COX15-related Leigh syndrome. Am. J. Med. Genet. A 2020, 182, 1506–1512.
  31. van Bon, B.W.M.; Oortveld, M.A.W.; Nijtmans, L.G.; Fenckova, M.; Nijhof, B.; Besseling, J.; Vos, M.; Kramer, J.M.; de Leeuw, N.; Castells-Nobau, A.; et al. CEP89 is required for mitochondrial metabolism and neuronal function in man and fly. Hum. Mol. Genet. 2013, 22, 3138–3151.
  32. Weraarpachai, W.; Sasarman, F.; Nishimura, T.; Antonicka, H.; Auré, K.; Rötig, A.; Lombès, A.; Shoubridge, E.A. Mutations in C12orf62, a factor that couples COX I synthesis with cytochrome c oxidase assembly, cause fatal neonatal lactic acidosis. Am. J. Hum. Genet. 2012, 90, 142–151.
  33. Ostergaard, E.; Weraarpachai, W.; Ravn, K.; Born, A.P.; Jønson, L.; Duno, M.; Wibrand, F.; Shoubridge, E.A.; Vissing, J. Mutations in COA3 cause isolated complex IV deficiency associated with neuropathy, exercise intolerance, obesity, and short stature. J. Med. Genet. 2015, 52, 203–207.
  34. Huigsloot, M.; Nijtmans, L.G.; Szklarczyk, R.; Baars, M.J.H.; van den Brand, M.A.M.; Hendriksfranssen, M.G.M.; van den Heuvel, L.P.; Smeitink, J.A.M.; Huynen, M.A.; Rodenburg, R.J.T. A mutation in C2orf64 causes impaired cytochrome c oxidase assembly and mitochondrial cardiomyopathy. Am. J. Hum. Genet. 2011, 88, 488–493.
  35. Martinez Lyons, A.; Ardissone, A.; Reyes, A.; Robinson, A.J.; Moroni, I.; Ghezzi, D.; Fernandez-Vizarra, E.; Zeviani, M. COA7 (C1orf163/RESA1) mutations associated with mitochondrial leukoencephalopathy and cytochrome c oxidase deficiency. J. Med. Genet. 2016, 53, 846–849.
  36. Higuchi, Y.; Okunushi, R.; Hara, T.; Hashiguchi, A.; Yuan, J.; Yoshimura, A.; Murayama, K.; Ohtake, A.; Ando, M.; Hiramatsu, Y.; et al. Mutations in COA7 cause spinocerebellar ataxia with axonal neuropathy. Brain 2018, 141, 1622–1636.
  37. Mohanraj, K.; Wasilewski, M.; Benincá, C.; Cysewski, D.; Poznanski, J.; Sakowska, P.; Bugajska, Z.; Deckers, M.; Dennerlein, S.; Fernandez-Vizarra, E.; et al. Inhibition of proteasome rescues a pathogenic variant of respiratory chain assembly factor COA7. EMBO Mol. Med. 2019, 11.
  38. Mansour, H.; Sabbagh, S.; Bizzari, S.; El-Hayek, S.; Chouery, E.; Gambarini, A.; Gencik, M.; Mégarbané, A. The Lebanese Allele in the PET100 Gene: Report on Two New Families with Cytochrome c Oxidase Deficiency. J. Pediatr. Genet. 2019, 8, 172–178.
  39. Oláhová, M.; Haack, T.B.; Alston, C.L.; Houghton, J.A.; He, L.; Morris, A.A.; Brown, G.K.; McFarland, R.; Chrzanowska-Lightowlers, Z.M.; Lightowlers, R.N.; et al. A truncating PET100 variant causing fatal infantile lactic acidosis and isolated cytochrome c oxidase deficiency. Eur. J. Hum. Genet. 2015, 23, 935–939.
  40. Lim, S.C.; Smith, K.R.; Stroud, D.A.; Compton, A.G.; Tucker, E.J.; Dasvarma, A.; Gandolfo, L.C.; Marum, J.E.; McKenzie, M.; Peters, H.L.; et al. A founder mutation in PET100 causes isolated complex IV deficiency in Lebanese individuals with Leigh syndrome. Am. J. Hum. Genet. 2014, 94, 209–222.
  41. Renkema, G.H.; Visser, G.; Baertling, F.; Wintjes, L.T.; Wolters, V.M.; van Montfrans, J.; de Kort, G.A.P.; Nikkels, P.G.J.; van Hasselt, P.M.; van der Crabben, S.N.; et al. Mutated PET117 causes complex IV deficiency and is associated with neurodevelopmental regression and medulla oblongata lesions. Hum. Genet. 2017, 136, 759–769.
  42. Antonicka, H.; Leary, S.C.; Guercin, G.-H.; Agar, J.N.; Horvath, R.; Kennaway, N.G.; Harding, C.O.; Jaksch, M.; Shoubridge, E.A. Mutations in COX10 result in a defect in mitochondrial heme A biosynthesis and account for multiple, early-onset clinical phenotypes associated with isolated COX deficiency. Hum. Mol. Genet. 2003, 12, 2693–2702.
  43. Coenen, M.J.H.; van den Heuvel, L.P.; Ugalde, C.; Ten Brinke, M.; Nijtmans, L.G.J.; Trijbels, F.J.M.; Beblo, S.; Maier, E.M.; Muntau, A.C.; Smeitink, J.A.M. Cytochrome c oxidase biogenesis in a patient with a mutation in COX10 gene. Ann. Neurol. 2004, 56, 560–564.
  44. Antonicka, H.; Mattman, A.; Carlson, C.G.; Glerum, D.M.; Hoffbuhr, K.C.; Leary, S.C.; Kennaway, N.G.; Shoubridge, E.A. Mutations in COX15 produce a defect in the mitochondrial heme biosynthetic pathway, causing early-onset fatal hypertrophic cardiomyopathy. Am. J. Hum. Genet. 2003, 72, 101–114.
  45. Oquendo, C.E.; Antonicka, H.; Shoubridge, E.A.; Reardon, W.; Brown, G.K. Functional and genetic studies demonstrate that mutation in the COX15 gene can cause Leigh syndrome. J. Med. Genet. 2004, 41, 540–544.
  46. Alfadhel, M.; Lillquist, Y.P.; Waters, P.J.; Sinclair, G.; Struys, E.; McFadden, D.; Hendson, G.; Hyams, L.; Shoffner, J.; Vallance, H.D. Infantile cardioencephalopathy due to a COX15 gene defect: Report and review. Am. J. Med. Genet. A 2011, 155A, 840–844.
  47. Bugiani, M.; Tiranti, V.; Farina, L.; Uziel, G.; Zeviani, M. Novel mutations in COX15 in a long surviving Leigh syndrome patient with cytochrome c oxidase deficiency. J. Med. Genet. 2005, 42, e28.
  48. Papadopoulou, L.C.; Sue, C.M.; Davidson, M.M.; Tanji, K.; Nishino, I.; Sadlock, J.E.; Krishna, S.; Walker, W.; Selby, J.; Glerum, D.M.; et al. Fatal infantile cardioencephalomyopathy with COX deficiency and mutations in SCO2, a COX assembly gene. Nat. Genet. 1999, 23, 333–337.
  49. Jaksch, M.; Ogilvie, I.; Yao, J.; Kortenhaus, G.; Bresser, H.G.; Gerbitz, K.D.; Shoubridge, E.A. Mutations in SCO2 are associated with a distinct form of hypertrophic cardiomyopathy and cytochrome c oxidase deficiency. Hum. Mol. Genet. 2000, 9, 795–801.
  50. Rebelo, A.P.; Saade, D.; Pereira, C.P.; Farooq, A.; Huff, T.C.; Abreu, L.; Moraes, C.T.; Mnatsakanova, D.; Mathews, K.; Yang, H.; et al. SCO2 mutations cause early-onset axonal Charcot-Marie-Tooth disease associated with cellular copper deficiency. Brain 2018, 141, 662–672.
  51. Jaksch, M.; Horvath, R.; Horn, N.; Auer, D.P.; Macmillan, C.; Peters, J.; Gerbitz, K.D.; Kraegeloh-Mann, I.; Muntau, A.; Karcagi, V.; et al. Homozygosity (E140K) in SCO2 causes delayed infantile onset of cardiomyopathy and neuropathy. Neurology 2001, 57, 1440–1446.
  52. Valnot, I.; Osmond, S.; Gigarel, N.; Mehaye, B.; Amiel, J.; Cormier-Daire, V.; Munnich, A.; Bonnefont, J.P.; Rustin, P.; Rötig, A. Mutations of the SCO1 gene in mitochondrial cytochrome c oxidase deficiency with neonatal-onset hepatic failure and encephalopathy. Am. J. Hum. Genet. 2000, 67, 1104–1109.
  53. Banci, L.; Bertini, I.; Ciofi-Baffoni, S.; Leontari, I.; Martinelli, M.; Palumaa, P.; Sillard, R.; Wang, S. Human Sco1 functional studies and pathological implications of the P174L mutant. Proc. Natl. Acad. Sci. USA 2007, 104, 15–20.
  54. Melchionda, L.; Haack, T.B.; Hardy, S.; Abbink, T.E.M.; Fernandez-Vizarra, E.; Lamantea, E.; Marchet, S.; Morandi, L.; Moggio, M.; Carrozzo, R.; et al. Mutations in APOPT1, encoding a mitochondrial protein, cause cavitating leukoencephalopathy with cytochrome c oxidase deficiency. Am. J. Hum. Genet. 2014, 95, 315–325.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 443
Revisions: 2 times (View History)
Update Date: 10 Oct 2020
1000/1000