How mitochondria produce reactive oxygen species

Michael P Murphy, Michael P Murphy

Abstract

The production of ROS (reactive oxygen species) by mammalian mitochondria is important because it underlies oxidative damage in many pathologies and contributes to retrograde redox signalling from the organelle to the cytosol and nucleus. Superoxide (O2(*-)) is the proximal mitochondrial ROS, and in the present review I outline the principles that govern O2(*-) production within the matrix of mammalian mitochondria. The flux of O2(*-) is related to the concentration of potential electron donors, the local concentration of O2 and the second-order rate constants for the reactions between them. Two modes of operation by isolated mitochondria result in significant O2(*-) production, predominantly from complex I: (i) when the mitochondria are not making ATP and consequently have a high Deltap (protonmotive force) and a reduced CoQ (coenzyme Q) pool; and (ii) when there is a high NADH/NAD+ ratio in the mitochondrial matrix. For mitochondria that are actively making ATP, and consequently have a lower Deltap and NADH/NAD+ ratio, the extent of O2(*-) production is far lower. The generation of O2(*-) within the mitochondrial matrix depends critically on Deltap, the NADH/NAD+ and CoQH2/CoQ ratios and the local O2 concentration, which are all highly variable and difficult to measure in vivo. Consequently, it is not possible to estimate O2(*-) generation by mitochondria in vivo from O2(*-)-production rates by isolated mitochondria, and such extrapolations in the literature are misleading. Even so, the description outlined here facilitates the understanding of factors that favour mitochondrial ROS production. There is a clear need to develop better methods to measure mitochondrial O2(*-) and H2O2 formation in vivo, as uncertainty about these values hampers studies on the role of mitochondrial ROS in pathological oxidative damage and redox signalling.

Figures

Figure 1. Overview of mitochondrial ROS production
Figure 1. Overview of mitochondrial ROS production
ROS production by mitochondria can lead to oxidative damage to mitochondrial proteins, membranes and DNA, impairing the ability of mitochondria to synthesize ATP and to carry out their wide range of metabolic functions, including the tricarboxylic acid cycle, fatty acid oxidation, the urea cycle, amino acid metabolism, haem synthesis and FeS centre assembly that are central to the normal operation of most cells. Mitochondrial oxidative damage can also increase the tendency of mitochondria to release intermembrane space proteins such as cytochrome c (cyt c) to the cytosol by mitochondrial outer membrane permeabilization (MOMP) and thereby activate the cell's apoptotic machinery. In addition, mitochondrial ROS production leads to induction of the mitochondrial permeability transition pore (PTP), which renders the inner membrane permeable to small molecules in situations such as ischaemia/reperfusion injury. Consequently, it is unsurprising that mitochondrial oxidative damage contributes to a wide range of pathologies. In addition, mitochondrial ROS may act as a modulatable redox signal, reversibly affecting the activity of a range of functions in the mitochondria, cytosol and nucleus.
Figure 2. Measurement of H 2 O…
Figure 2. Measurement of H2O2 production by isolated mitochondria
The production of O2•− within the mitochondrial matrix, intermembrane space and outer membrane leads to the formation of H2O2 from SOD-catalysed dismutation. Some O2•− can react directly with nitric oxide (NO•) to form peroxynitrite (ONOO−). There are also sources outside mitochondria that produce H2O2 directly. The H2O2 efflux from mitochondria can be measured following reaction with a non-fluorescent substrate such as Amplex Red in conjunction with horseradish peroxidase (HRP) to form a fluorescent product, resorufin. Within mitochondria H2O2 is degraded by glutathione peroxidases (GPx) or peroxiredoxins (Prx) which depend on glutathione (GSH) and thioredoxin-2 (Trx) for their reduction respectively. Glutathione disulfide (GSSG) is reduced back to GSH by glutathione reductase (GR). Trx is reduced by thioredoxin reductase-2 (TrxR). Both enzymes receive reducing equivalents from the NADPH pool, which is kept reduced by the Δp-dependent transhydrogenase (TH), and by isocitrate dehydrogenase (ICDH). Note that many mitochondrial preparations, particularly those from the liver, contain large amounts of catalase contamination. The effects of such extramitochondrial H2O2 sinks are not indicated here as they are usually accounted for by appropriate H2O2 calibration curves in the presence of mitochondria. ox, oxidized; red, reduced. An animated version of this Figure can be seen at http://www.BiochemJ.org/bj/417/0001/bj4170001add.htm.
Figure 3. Modes of mitochondrial operation that…
Figure 3. Modes of mitochondrial operation that lead to O2•− production
There are three modes of mitochondrial operation that are associated with O2•− production. In mode 1, the NADH pool is reduced, for example by damage to the respiratory chain, loss of cytochrome c during apoptosis or low ATP demand. This leads to a rate of O2•− formation at the FMN of complex I that is determined by the extent of FMN reduction which is in turn set by the NADH/NAD+ ratio. Other sites such as αKGDH may also contribute. In mode 2, there is no ATP production and there is a high Δp and a reduced CoQ pool which leads to RET through complex I, producing large amounts of O2•−. In mode 3, mitochondria are actively making ATP and consequently have a lower Δp than in mode 2 and a more oxidized NADH pool than in mode 1. Under these conditions, the flux of O2•− within mitochondria is far lower than in modes 1 and 2, and the O2•− sources are unclear.
Figure 4. Production of O 2 •−…
Figure 4. Production of O2•− by complex I
The cartoon of complex I is a chimaera modelled on the hydrophobic arm of Yarrowia lipolytica obtained by electron microscopy [132] and the crystal structure of the hydrophilic arm from Thermus thermophilus [82]. The location of the FMN and the FeS centres in the water-soluble arm are indicated, along with the putative CoQ-binding site. In mode 1, there is extensive O2•− production from the FMN in response to a reduced NADH pool. In mode 2, a high Δp and a reduced CoQ pool lead to RET and a high flux of O2•− from the complex. The site of this O2•− production is uncertain, hence the question mark, but may be associated with the CoQ-binding site(s).
Figure 5. Possible mechanisms of mitochondrial redox…
Figure 5. Possible mechanisms of mitochondrial redox signalling
The production of H2O2 from mitochondria is a potential redox signal. H2O2 generated by mitochondria can reversibly alter the activity of proteins with critical protein thiols by modifying them to intra- or inter-protein disulfides, or to mixed disulfides with GSH. These modifications can occur on mitochondrial, cytosolic or nuclear enzymes, carriers or transcription factors, transiently altering their activities. The change in activity can be reversed by reducing the modified protein thiol by endogenous thiol reductants such as GSH or thioredoxin. As the extent of H2O2 production from mitochondria will depend on factors such as Δp or the redox state of the NADH pool, it can act as a retrograde signal to the rest of the cell, reporting on mitochondrial status. This signal can then lead to the short-term modification of, for example, pathways supplying substrates to the mitochondria. Alternatively, longer-term modifications can occur through modifying redox-sensitive transcription factors that adjust the production of mitochondrial components. In addition, external signals may modify O2•− production by the respiratory chain by post-translational modification. Alteration of the activity of mitochondrial peroxidases could also modulate H2O2 efflux from mitochondria to the rest of the cell. It is also possible that secondary redox signals, such as lipid peroxidation products derived from H2O2, could act as secondary redox signals.

References

    1. Andreyev A. Y., Kushnareva Y. E., Starkov A. A. Mitochondrial metabolism of reactive oxygen species. Biochemistry (Moscow) 2005;70:200–214.
    1. Turrens J. F. Mitochondrial formation of reactive oxygen species. J. Physiol. 2003;552:335–344.
    1. Balaban R. S., Nemoto S., Finkel T. Mitochondria, oxidants, and aging. Cell. 2005;120:483–495.
    1. Chance B., Sies H., Boveris A. Hydroperoxide metabolism in mammalian organs. Physiol. Rev. 1979;59:527–605.
    1. Cadenas E., Davies K. J. Mitochondrial free radical generation, oxidative stress, and aging. Free Radical Biol. Med. 2000;29:222–230.
    1. Raha S., Robinson B. H. Mitochondria, oxygen free radicals, disease and ageing. Trends Biochem. 2000;25:502–508.
    1. Adam-Vizi V., Chinopoulos C. Bioenergetics and the formation of mitochondrial reactive oxygen species. Trends Pharmacol. Sci. 2006;27:639–645.
    1. Muller F. The nature and mechanism of superoxide production by the electron transport chain. J. Am. Aging Assoc. 2000;23:227–253.
    1. Droge W. Free radicals in the physiological control of cell function. Physiol. Rev. 2002;82:47–95.
    1. Jensen P. K. Antimycin-insensitive oxidation of succinate and reduced nicotinamide-adenine dinucleotide in electron-transport particles. I. pH dependency and hydrogen peroxide formation. Biochim. Biophys. Acta. 1966;122:157–166.
    1. Loschen G., Flohe L., Chance B. Respiratory chain linked H2O2 production in pigeon heart mitochondria. FEBS Lett. 1971;18:261–264.
    1. Boveris A., Chance B. The mitochondrial generation of hydrogen peroxide: general properties and effect of hyperbaric oxygen. Biochem. J. 1973;134:707–716.
    1. Loschen G., Azzi A., Richter C., Flohe L. Superoxide radicals as precursors of mitochondrial hydrogen peroxide. FEBS Lett. 1974;42:68–72.
    1. Forman H. J., Kennedy J. A. Role of superoxide radical in mitochondrial dehydrogenase reactions. Biochem. Biophys. Res. Commun. 1974;60:1044–1050.
    1. Weisiger R. A., Fridovich I. Superoxide dismutase: organelle specificity. J. Biol. Chem. 1973;248:3582–3592.
    1. Sawyer D. T., Valentine J. S. How super is superoxide? Acc. Chem. Res. 1981;14:393–400.
    1. Reynafarje B., Costa L. E., Lehninger A. L. O2 solubility in aqueous media determined by a kinetic method. Anal. Biochem. 1985;145:406–418.
    1. Erecinska M., Silver I. A. Tissue oxygen tension and brain sensitivity to hypoxia. Respir. Physiol. 2001;128:263–276.
    1. Skulachev V. P. Role of uncoupled and non-coupled oxidations in maintenance of safely low levels of oxygen and its one-electron reductants. Q. Rev. Biophys. 1996;29:169–202.
    1. Tyler D. D. Polarographic assay and intracellular distribution of superoxide dismutase in rat liver. Biochem. J. 1975;147:493–504.
    1. Winterbourn C. C., French J. K., Claridge R. F. Superoxide dismutase as an inhibitor of reactions of semiquinone radicals. FEBS Lett. 1978;94:269–272.
    1. Schafer F. Q., Buettner G. R. Redox environment of the cell as viewed through the redox state of the glutathione disulfide/glutathione couple. Free Radical Biol. Med. 2001;30:1191–1212.
    1. Barja G. Mitochondrial oxygen radical generation and leak: sites of production in states 4 and 3, organ specificity, and relation to aging and longevity. J. Bioenerg. Biomembr. 1999;31:347–366.
    1. St-Pierre J., Buckingham J. A., Roebuck S. J., Brand M. D. Topology of superoxide production from different sites in the mitochondrial electron transport chain. J. Biol. Chem. 2002;277:44784–44790.
    1. Wikstrom M., Krab K., Saraste M. London: Academic Press; 1981. Cytochrome Oxidase: a Synthesis.
    1. Turrens J. F., Freeman B. A., Levitt J. G., Crapo J. D. The effect of hyperoxia on superoxide production by lung submitochondrial particles. Arch. Biochem. Biophys. 1982;217:401–410.
    1. Kudin A. P., Bimpong-Buta N. Y., Vielhaber S., Elger C. E., Kunz W. S. Characterization of superoxide-producing sites in isolated brain mitochondria. J. Biol. Chem. 2004;279:4127–4135.
    1. Kussmaul L., Hirst J. The mechanism of superoxide production by NADH:ubiquinone oxidoreductase (complex I) from bovine heart mitochondria. Proc. Natl. Acad. Sci. U.S.A. 2006;103:7607–7612.
    1. Hoffman D. L., Salter J. D., Brookes P. S. Response of mitochondrial reactive oxygen species generation to steady-state oxygen tension: implications for hypoxic cell signaling. Am. J. Physiol. Heart Circ. Physiol. 2007;292:H101–H108.
    1. Alvarez S., Valdez L. B., Zaobornyj T., Boveris A. Oxygen dependence of mitochondrial nitric oxide synthase activity. Biochem. Biophys. Res. Commun. 2003;305:771–775.
    1. Shiva S., Brookes P. S., Patel R. P., Anderson P. G., Darley-Usmar V. M. Nitric oxide partitioning into mitochondrial membranes and the control of respiration at cytochrome c oxidase. Proc. Natl. Acad. Sci. U.S.A. 2001;98:7212–7217.
    1. Brown G. C., Cooper C. E. Nanomolar concentrations of nitric oxide reversibly inhibit synaptosomal respiration by competing with oxygen at cytochrome oxidase. FEBS Lett. 1994;356:295–298.
    1. Brown G. C. Nitric oxide regulates mitochondrial respiration and cell functions by inhibiting cytochrome oxidase. FEBS Lett. 1995;369:136–139.
    1. Cleeter M. J. W., Cooper J. M., Darley-Usmar V. M., Moncada S., Schapira A. H. V. Reversible inhibition of cytochrome c oxidase, the terminal enzyme of mitochondrial respiratory chain, by nitric oxide. FEBS Lett. 1994;345:50–54.
    1. Brown G. C. Nitric oxide and mitochondrial respiration. Biochim. Biophys. Acta. 1999;1411:351–369.
    1. Mateo J., García-Lecea M., Cadenas S., Hernández C., Moncada S. Regulation of hypoxia-inducible factor-1α by nitric oxide through mitochondria-dependent and -independent pathways. Biochem J. 2003;376:537–544.
    1. Hagen T., Taylor C. T., Lam F., Moncada S. Redistribution of intracellular oxygen in hypoxia by nitric oxide: effect on HIF1α. Science. 2003;302:1975–1978.
    1. Moncada S., Erusalimsky J. D. Does nitric oxide modulate mitochondrial energy generation and apoptosis? Nat. Rev. Mol. Cell Biol. 2002;3:214–220.
    1. Klinman J. P. How do enzymes activate oxygen without inactivating themselves? Acc. Chem. Res. 2007;40:325–333.
    1. Marcus R. A., Sutin N. Electron transfers in chemistry and biology. Biochim. Biophys. Acta. 1985;811:265–322.
    1. Moser C. C., Farid T. A., Chobot S. E., Dutton P. L. Electron tunneling chains of mitochondria. Biochim. Biophys. Acta. 2006;1757:1096–1109.
    1. Massey V. Activation of molecular oxygen by flavins and flavoproteins. J. Biol. Chem. 1994;269:22459–22462.
    1. Imlay J. A. A metabolic enzyme that rapidly produces superoxide, fumarate reductase of Escherichia coli. J. Biol. Chem. 1995;270:19767–19777.
    1. Cochemé H. M., Murphy M. P. Complex I is the major site of mitochondrial superoxide production by paraquat. J. Biol. Chem. 2008;283:1786–1798.
    1. Li Y., Huang T.-T., Carlson E. J., Melov S., Ursell P. C., Olson J. L., Noble L. J., Yoshimura M. P., Berger C., Chan P. H., et al. Dilated cardiomyopathy and neonatal lethality in mutant mice lacking manganese superoxide dismutase. Nat. Genet. 1995;11:376–381.
    1. Lebovitz R. M., Zhang H., Vogel H., Cartwright J., Dionne L., Lu N., Huang S., Matzuk M. M. Neurodegeneration, myocardial injury, and perinatal death in mitochondrial superoxide dismutase-deficient mice. Proc. Natl. Acad. Sci. U.S.A. 1996;93:9782–9787.
    1. Zhao H., Joseph J., Fales H. M., Sokoloski E. A., Levine R. L., Vasquez-Vivar J., Kalyanaraman B. Detection and characterization of the product of hydroethidine and intracellular superoxide by HPLC and limitations of fluorescence. Proc. Natl. Acad. Sci. U.S.A. 2005;102:5727–5732.
    1. Robinson K. M., Janes M. S., Pehar M., Monette J. S., Ross M. F., Hagen T. M., Murphy M. P., Beckman J. S. Selective fluorescent imaging of superoxide in vivo using ethidium-based probes. Proc. Natl. Acad. Sci. U.S.A. 2006;103:15038–15043.
    1. Zielonka J., Vasquez-Vivar J., Kalyanaraman B. Detection of 2-hydroxyethidium in cellular systems: a unique marker product of superoxide and hydroethidine. Nat. Protoc. 2008;3:8–21.
    1. Zielonka J., Srinivasan S., Hardy M., Ouari O., Lopez M., Vasquez-Vivar J., Avadhani N. G., Kalyanaraman B. Cytochrome c-mediated oxidation of hydroethidine and mito-hydroethidine in mitochondria: identification of homo- and heterodimers. Free Radical Biol. Med. 2008;44:835–846.
    1. Lucas M., Solano F. Coelenterazine is a superoxide anion-sensitive chemiluminescent probe: its usefulness in the assay of respiratory burst in neutrophils. Anal. Biochem. 1992;206:273–277.
    1. Han D., Antunes F., Canali R., Rettori D., Cadenas E. Voltage-dependent anion channels control the release of the superoxide anion from mitochondria to cytosol. J. Biol. Chem. 2003;278:5557–5563.
    1. Gardner P. R. Aconitase: sensitive target and measure of superoxide. Methods Enzmol. 2002;349:9–23.
    1. Boveris A. Determination of the production of superoxide radicals and hydrogen peroxide in mitochondria. Methods Enzymol. 1984;105:429–435.
    1. Barja G. Methods in Aging Research. Boca Raton: CRC Press; 1999. Kinetic measurement of mitochondrial oxygen radical production; pp. 533–548.
    1. Sohal R. S. Hydrogen peroxide production by mitochondria may be a biomarker of aging. Mech. Ageing Dev. 1991;60:189–198.
    1. Hansford R. G., Hogue B. A., Mildaziene V. Dependence of H2O2 formation by rat heart mitochondria on substrate availability and donor age. J. Bioenerg. Biomembr. 1997;29:89–95.
    1. Szabo C., Ischiropoulos H., Radi R. Peroxynitrite: biochemistry, pathophysiology and development of therapeutics. Nat. Rev. 2007;6:662–680.
    1. Packer M. A., Porteous C. M., Murphy M. P. Mitochondrial superoxide production in the presence of nitric oxide leads to the formation of peroxynitrite. Biochem. Mol. Biol. Int. 1996;40:527–534.
    1. Giorgio M., Migliaccio E., Orsini F., Paolucci D., Moroni M., Contursi C., Pelliccia G., Luzi L., Minucci S., Marcaccio M., et al. Electron transfer between cytochrome c and p66Shc generates reactive oxygen species that trigger mitochondrial apoptosis. Cell. 2005;122:221–233.
    1. Zoccarato F., Cavallini L., Alexandre A. Respiration-dependent removal of exogenous H2O2 in brain mitochondria: inhibition by Ca2+ J. Biol. Chem. 2004;279:4166–4174.
    1. Rhee S. G., Yang K. S., Kang S. W., Woo H. A., Chang T. S. Controlled elimination of intracellular H2O2: regulation of peroxiredoxin, catalase, and glutathione peroxidase via post-translational modification. Antioxid. Redox Signaling. 2005;7:619–626.
    1. Hurd T. R., Costa N. J., Dahm C. C., Beer S. M., Brown S. E., Filipovska A., Murphy M. P. Glutathionylation of mitochondrial proteins. Antioxid. Redox Signaling. 2005;7:999–1010.
    1. Rhee S. G., Kang S. W., Chang T. S., Jeong W., Kim K. Peroxiredoxin, a novel family of peroxidases. IUBMB Life. 2001;52:35–41.
    1. Salvi M., Battaglia V., Brunati A. M., La Rocca N., Tibaldi E., Pietrangeli P., Marcocci L., Mondovi B., Rossi C. A., Toninello A. Catalase takes part in rat liver mitochondria oxidative stress defense. J. Biol. Chem. 2007;282:24407–24415.
    1. Radi R., Turrens J. F., Chang L. Y., Bush K. M., Crapo J. D., Freeman B. A. Detection of catalase in rat heart mitochondria. J. Biol. Chem. 1991;266:22028–22034.
    1. Imai H., Nakagawa Y. Biological significance of phospholipid hydroperoxide glutathione peroxidase (PHGPx, GPx4) in mammalian cells. Free Radical Biol. Med. 2003;34:145–169.
    1. Rigobello M. P., Folda A., Scutari G., Bindoli A. The modulation of thiol redox state affects the production and metabolism of hydrogen peroxide by heart mitochondria. Arch. Biochem. Biophys. 2005;441:112–122.
    1. Rhee S. G., Kang S. W., Jeong W., Chang T. S., Yang K. S., Woo H. A. Intracellular messenger function of hydrogen peroxide and its regulation by peroxiredoxins. Curr. Opin. Cell Biol. 2005;17:183–189.
    1. Cox A. G., Pullar J. M., Hughes G., Ledgerwood E. C., Hampton M. B. Oxidation of mitochondrial peroxiredoxin 3 during the initiation of receptor-mediated apoptosis. Free Radical Biol. Med. 2008;44:1001–1009.
    1. Hurd T. R., Prime T. A., Harbour M. E., Lilley K. S., Murphy M. P. Detection of reactive oxygen species-sensitive thiol proteins by redox difference gel electrophoresis: implications for mitochondrial redox signaling. J. Biol. Chem. 2007;282:22040–22051.
    1. Holmgren A. Thioredoxin. Annu. Rev. Biochem. 1985;54:237–271.
    1. Hurd T. R., Filipovska A., Costa N. J., Dahm C. C., Murphy M. P. Disulphide formation on mitochondrial protein thiols. Biochem. Soc. Trans. 2005;33:1390–1393.
    1. Rydström J. Mitochondrial NADPH, transhydrogenase and disease. Biochim. Biophys. Acta. 2006;1757:721–726.
    1. Sazanov L. A., Jackson J. B. Proton-translocating transhydrogenase and NAD- and NADP-linked isocitrate dehydrogenases operate in a substrate cycle which contributes to fine regulation of the tricarboxylic acid cycle activity in mitochondria. FEBS Lett. 1994;344:109–116.
    1. Bienert G. P., Moller A. L., Kristiansen K. A., Schulz A., Moller I. M., Schjoerring J. K., Jahn T. P. Specific aquaporins facilitate the diffusion of hydrogen peroxide across membranes. J. Biol. Chem. 2007;282:1183–1192.
    1. Bienert G. P., Schjoerring J. K., Jahn T. P. Membrane transport of hydrogen peroxide. Biochim. Biophys. Acta. 2006;1758:994–1003.
    1. Korshunov S. S., Skulachev V. P., Starkov A. A. High protonic potential actuates a mechanism of production of reactive oxygen species in mitochondria. FEBS Lett. 1997;416:15–18.
    1. Lambert A. J., Brand M. D. Superoxide production by NADH:ubiquinone oxidoreductase (complex I) depends on the pH gradient across the mitochondrial inner membrane. Biochem. J. 2004;382:511–517.
    1. Hirst J., Carroll J., Fearnley I. M., Shannon R. J., Walker J. E. The nuclear encoded subunits of complex I from bovine heart mitochondria. Biochim. Biophys. Acta. 2003;1604:135–150.
    1. Sazanov L. A. Respiratory complex I: mechanistic and structural insights provided by the crystal structure of the hydrophilic domain. Biochemistry. 2007;46:2275–2288.
    1. Sazanov L. A., Hinchliffe P. Structure of the hydrophilic domain of respiratory complex I from Thermus thermophilus. Science. 2006;311:1430–1436.
    1. Hinkle P. C., Butow R. A., Racker E., Chance B. Partial resolution of the enzymes catalyzing oxidative phosphorylation. XV. Reverse electron transfer in the flavin-cytochrome β region of the respiratory chain of beef heart submitochondrial particles. J. Biol. Chem. 1967;242:5169–5173.
    1. Cadenas E., Boveris A., Ragan C. I., Stoppani A. O. Production of superoxide radicals and hydrogen peroxide by NADH–ubiquinone reductase and ubiquinol–cytochrome c reductase from beef-heart mitochondria. Arch. Biochem. Biophys. 1977;180:248–257.
    1. Hirst J., King M. S., Pryde K. R. The production of reactive oxygen species by complex I. Biochem. Soc. Trans. 2008;36:976–980.
    1. Takeshige K., Minakami S. NADH- and NADPH-dependent formation of superoxide anions by bovine heart submitochondrial particles and NADH–ubiquinone reductase preparation. Biochem. J. 1979;180:129–135.
    1. Votyakova T. V., Reynolds I. J. Δψm-Dependent and -independent production of reactive oxygen species by rat brain mitochondria. J. Neurochem. 2001;79:266–277.
    1. Kushnareva Y., Murphy A. N., Andreyev A. Complex I-mediated reactive oxygen species generation: modulation by cytochrome c and NAD(P)+ oxidation–reduction state. Biochem. J. 2002;368:545–553.
    1. Liu Y., Fiskum G., Schubert D. Generation of reactive oxygen species by the mitochondrial electron transport chain. J. Neurochem. 2002;80:780–787.
    1. Boveris A., Oshino N., Chance B. The cellular production of hydrogen peroxide. Biochem. J. 1972;128:617–630.
    1. Seo B. B., Marella M., Yagi T., Matsuno-Yagi A. The single subunit NADH dehydrogenase reduces generation of reactive oxygen species from complex I. FEBS Lett. 2006;580:6105–6108.
    1. Chance B., Hollunger G. The interaction of energy and electron transfer reactions in mitochondria. I. General properties and nature of the products of succinate-linked reduction of pyridine nucleotide. J Biol. Chem. 1961;236:1534–1543.
    1. Krishnamoorthy G., Hinkle P. C. Studies on the electron transfer pathway, topography of iron–sulfur centers, and site of coupling in NADH-Q oxidoreductase. J. Biol. Chem. 1988;263:17566–17575.
    1. Cino M., Del Maestro R. F. Generation of hydrogen peroxide by brain mitochondria: the effect of reoxygenation following postdecapitative ischemia. Arch. Biochem. Biophys. 1989;269:623–638.
    1. Lambert A. J., Brand M. D. Inhibitors of the quinone-binding site allow rapid superoxide production from mitochondrial NADH:ubiquinone oxidoreductase (complex I) J. Biol. Chem. 2004;279:39414–39420.
    1. Liu S. S. Generating, partitioning, targeting and functioning of superoxide in mitochondria. Biosci. Rep. 1997;17:259–272.
    1. Lambert A. J., Buckingham J. A., Boysen H. M., Brand M. D. Diphenyleneiodonium acutely inhibits reactive oxygen species production by mitochondrial complex I during reverse, but not forward electron transport. Biochim. Biophys. Acta. 2008;1777:397–403.
    1. Lambert A. J., Buckingham J. A., Brand M. D. Dissociation of superoxide production by mitochondrial complex I from NAD(P)H redox state. FEBS Lett. 2008;582:1711–1714.
    1. Iwata S., Lee J. W., Okada K., Lee J. K., Iwata M., Rasmussen B., Link T. A., Ramaswamy S., Jap B. K. Complete structure of the 11-subunit bovine mitochondrial cytochrome bc1 complex. Science. 1998;281:64–71.
    1. Turrens J. F., Alexandre A., Lehninger A. L. Ubisemiquinone is the electron donor for superoxide formation by complex III of heart mitochondria. Arch. Biochem. Biophys. 1985;237:408–414.
    1. Zhang L., Yu L., Yu C. A. Generation of superoxide anion by succinate–cytochrome c reductase from bovine heart mitochondria. J. Biol. Chem. 1998;273:33972–33976.
    1. Rich P. R., Bonner W. D. The sites of superoxide anion generation in higher plant mitochondria. Arch. Biochem. Biophys. 1978;188:206–213.
    1. Grigolava I. V., Ksenzenko M., Konstantinob A. A., Tikhonov A. N., Kerimov T. M. Tiron as a spin-trap for superoxide radicals produced by the respiratory chain of submitochondrial particles. Biochemisty (Moscow) 1980;45:75–82.
    1. Muller F. L., Liu Y., Van Remmen H. Complex III releases superoxide to both sides of the inner mitochondrial membrane. J. Biol. Chem. 2004;279:49064–49073.
    1. Forman H. J., Azzi A. On the virtual existence of superoxide anions in mitochondria: thoughts regarding its role in pathophysiology. FASEB J. 1997;11:374–375.
    1. Starkov A. A., Fiskum G., Chinopoulos C., Lorenzo B. J., Browne S. E., Patel M. S., Beal M. F. Mitochondrial α-ketoglutarate dehydrogenase complex generates reactive oxygen species. J. Neurosci. 2004;24:7779–7788.
    1. Tretter L., Adam-Vizi V. Generation of reactive oxygen species in the reaction catalyzed by α-ketoglutarate dehydrogenase. J. Neurosci. 2004;24:7771–7778.
    1. Bunik V. I., Sievers C. Inactivation of the 2-oxo acid dehydrogenase complexes upon generation of intrinsic radical species. Eur. J. Biochem. 2002;269:5004–5015.
    1. Eaton S. Control of mitochondrial β-oxidation flux. Prog. Lipid Res. 2002;41:197–239.
    1. Forman H. J., Kennedy J. Dihydroorotate-dependent superoxide production in rat brain and liver: a function of the primary dehydrogenase. Arch. Biochem. Biophys. 1976;173:219–224.
    1. Koza R. A., Kozak U. C., Brown L. J., Leiter E. H., MacDonald M. J., Kozak L. P. Sequence and tissue-dependent RNA expression of mouse FAD-linked glycerol-3-phosphate dehydrogenase. Arch. Biochem. Biophys. 1996;336:97–104.
    1. Drahota Z., Chowdhury S. K., Floryk D., Mracek T., Wilhelm J., Rauchova H., Lenaz G., Houstek J. Glycerophosphate-dependent hydrogen peroxide production by brown adipose tissue mitochondria and its activation by ferricyanide. J. Bioenerg. Biomembr. 2002;34:105–113.
    1. Tretter L., Takacs K., Hegedus V., Adam-Vizi V. Characteristics of α-glycerophosphate-evoked H2O2 generation in brain mitochondria. J. Neurochem. 2007;100:650–663.
    1. Miwa S., St-Pierre J., Partridge L., Brand M. D. Superoxide and hydrogen peroxide production by Drosophila mitochondria. Free Radical Biol. Med. 2003;35:938–948.
    1. Guzy R. D., Sharma B., Bell E., Chandel N. S., Schumacker P. T. Loss of the SdhB, but not the SdhA, subunit of complex II triggers reactive oxygen species-dependent hypoxia-inducible factor activation and tumorigenesis. Mol. Cell. Biol. 2008;28:718–731.
    1. Hanukoglu I. Antioxidant protective mechanisms against reactive oxygen species (ROS) generated by mitochondrial P450 systems in steroidogenic cells. Drug Metab. Rev. 2006;38:171–196.
    1. Hanukoglu I., Rapoport R., Weiner L., Sklan D. Electron leakage from the mitochondrial NADPH–adrenodoxin reductase–adrenodoxin–P450scc (cholesterol side chain cleavage) system. Arch. Biochem. Biophys. 1993;305:489–498.
    1. Brand M. D., Affourtit C., Esteves T. C., Green K., Lambert A. J., Miwa S., Pakay J. L., Parker N. Mitochondrial superoxide: production, biological effects, and activation of uncoupling proteins. Free Radical Biol. Med. 2004;37:755–767.
    1. Starkov A. A., Fiskum G. Regulation of brain mitochondrial H2O2 production by membrane potential and NAD(P)H redox state. J. Neurochem. 2003;86:1101–1107.
    1. Chandel N. S., Maltepe E., Goldwasser E., Mathieu C. E., Simon M. C., Schumacker P. T. Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc. Natl. Acad. Sci. U.S.A. 1998;95:11715–11720.
    1. Guzy R. D., Schumacker P. T. Oxygen sensing by mitochondria at complex III: the paradox of increased reactive oxygen species during hypoxia. Exp. Physiol. 2006;91:807–819.
    1. Schofield C. J., Ratcliffe P. J. Oxygen sensing by HIF hydroxylases. Nat. Rev. Mol. Cell Biol. 2004;5:343–354.
    1. Semenza G. L. O2-regulated gene expression: transcriptional control of cardiorespiratory physiology by HIF-1. J. Appl. Physiol. 2004;96:1173–1177.
    1. Sanjuan-Pla A., Cervera A. M., Apostolova N., Garcia-Bou R., Victor V. M., Murphy M. P., McCreath K. J. A targeted antioxidant reveals the importance of mitochondrial reactive oxygen species in the hypoxic signaling of HIF-1α. FEBS Lett. 2005;579:2669–2674.
    1. Guzy R. D., Hoyos B., Robin E., Chen H., Liu L., Mansfield K. D., Simon M. C., Hammerling U., Schumacker P. T. Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metab. 2005;1:401–408.
    1. Bell E. L., Klimova T. A., Eisenbart J., Moraes C. T., Murphy M. P., Budinger G. R., Chandel N. S. The Qo site of the mitochondrial complex III is required for the transduction of hypoxic signaling via reactive oxygen species production. J. Cell Biol. 2007;177:1029–1036.
    1. Palacios-Callender M., Quintero M., Hollis V. S., Springett R. J., Moncada S. Endogenous NO regulates superoxide production at low oxygen concentrations by modifying the redox state of cytochrome c oxidase. Proc. Natl. Acad. Sci. U.S.A. 2004;101:7630–7635.
    1. Beckman K. B., Ames B. N. The free radical theory of aging matures. Physiol. Rev. 1998;78:547–581.
    1. Oshino N., Jamieson D., Chance B. The properties of hydrogen peroxide production under hyperoxic and hypoxic conditions of perfused rat liver. Biochem. J. 1975;146:53–65.
    1. Murphy M. P., Smith R. A. Targeting antioxidants to mitochondria by conjugation to lipophilic cations. Annu. Rev. Pharmacol. Toxicol. 2007;47:629–656.
    1. Hurd T. R., Prime T. A., Harbour M. E., Lilley K. S., Murphy M. P. Detection of reactive oxygen species-sensitive thiol proteins by redox difference gel electrophoresis: implications for mitochondrial redox signaling. J. Biol. Chem. 2007;282:22040–22051.
    1. Radermacher M., Ruiz T., Clason T., Benjamin S., Brandt U., Zickermann V. The three-dimensional structure of complex I from Yarrowia lipolytica: a highly dynamic enzyme. J. Struct. Biol. 2006;154:269–279.

Source: PubMed

3
Subskrybuj