6,036
Views
50
CrossRef citations to date
0
Altmetric
Review Articles

Riding the tiger – physiological and pathological effects of superoxide and hydrogen peroxide generated in the mitochondrial matrix

Pages 592-661 | Received 16 Jun 2020, Accepted 22 Sep 2020, Published online: 04 Nov 2020

Abstract

Elevated mitochondrial matrix superoxide and/or hydrogen peroxide concentrations drive a wide range of physiological responses and pathologies. Concentrations of superoxide and hydrogen peroxide in the mitochondrial matrix are set mainly by rates of production, the activities of superoxide dismutase-2 (SOD2) and peroxiredoxin-3 (PRDX3), and by diffusion of hydrogen peroxide to the cytosol. These considerations can be used to generate criteria for assessing whether changes in matrix superoxide or hydrogen peroxide are both necessary and sufficient to drive redox signaling and pathology: is a phenotype affected by suppressing superoxide and hydrogen peroxide production; by manipulating the levels of SOD2, PRDX3 or mitochondria-targeted catalase; and by adding mitochondria-targeted SOD/catalase mimetics or mitochondria-targeted antioxidants? Is the pathology associated with variants in SOD2 and PRDX3 genes? Filtering the large literature on mitochondrial redox signaling using these criteria highlights considerable evidence that mitochondrial superoxide and hydrogen peroxide drive physiological responses involved in cellular stress management, including apoptosis, autophagy, propagation of endoplasmic reticulum stress, cellular senescence, HIF1α signaling, and immune responses. They also affect cell proliferation, migration, differentiation, and the cell cycle. Filtering the huge literature on pathologies highlights strong experimental evidence that 30-40 pathologies may be driven by mitochondrial matrix superoxide or hydrogen peroxide. These can be grouped into overlapping and interacting categories: metabolic, cardiovascular, inflammatory, and neurological diseases; cancer; ischemia/reperfusion injury; aging and its diseases; external insults, and genetic diseases. Understanding the involvement of mitochondrial matrix superoxide and hydrogen peroxide concentrations in these diseases can facilitate the rational development of appropriate therapies.

Introduction

Aerobic life must deal with a dilemma. The most abundant and accessible source of energy on Earth is sunlight, and that energy is stored by plants primarily as reduced carbon (–CH2–, –CHOH–, etc.). The best way to release the energy is to pass electrons from the reduced carbon substrates to the most oxidizing acceptor that is widely available, oxygen. However, electrons are very reactive and hard to control, so riding the tiger of abundant energy release carries the risk of premature electron escape to oxygen to form reactive oxygen species (ROS) that can drive damage and disease – the tiger may bite you.

Mitochondria take electrons from reduced carbon substrates such as fatty acids, amino acids and tricarboxylic acid cycle intermediates using appropriate substrate dehydrogenases. They pass these electrons down an electron transport chain, and use them two pairs at a time to reduce molecular oxygen (O2) to water (2H2O). The redox energy that is released by substrate oxidation is coupled to proton pumping from the mitochondrial matrix to the intermembrane space, and the resulting protonmotive force across the mitochondrial inner membrane, consisting of a membrane potential and a pH gradient, is used to drive ATP synthesis during oxidative phosphorylation. Occasionally, however, electrons leak prematurely from the respiratory chain to reduce O2 singly to form the superoxide radical anion (O2•−), or in pairs to form hydrogen peroxide (H2O2) (Boveris and Chance Citation1973; Chance et al. Citation1979). Superoxide and hydrogen peroxide are the two primary ROS from which others, such as hydroxyl radical and peroxynitrite, are formed.

There is a large literature on the roles of mitochondrially-generated superoxide and hydrogen peroxide in cellular signaling (Cadenas Citation2004; Starkov Citation2008; Hamanaka and Chandel Citation2010; Ray et al. Citation2012; Sena and Chandel Citation2012; Guo et al. Citation2013; Dai et al. Citation2014; Holmstrom and Finkel Citation2014; Shadel and Horvath Citation2015; Brand Citation2016; Diebold and Chandel Citation2016; Siauciunaite et al. Citation2019; Watson et al. Citation2019; Sies and Jones Citation2020). An even larger literature invokes mitochondrial ROS production as a cause of oxidative damage and pathology (Cadenas and Davies Citation2000; Raha and Robinson Citation2000; Lenaz Citation2001; Turrens Citation2003; Brand et al. Citation2004; Brookes et al. Citation2004; Cadenas Citation2004; Adam-Vizi Citation2005; Andreyev et al. Citation2005; Balaban et al. Citation2005; Brookes Citation2005; Jezek and Hlavata Citation2005; Zorov et al. Citation2006; Starkov Citation2008; Kowaltowski et al. Citation2009; Lambert and Brand Citation2009; Murphy Citation2009; Brand Citation2010; Jastroch et al. Citation2010; Rigoulet et al. Citation2011; Drose and Brandt Citation2012; Sena and Chandel Citation2012; Chen and Zweier Citation2014; Holmstrom and Finkel Citation2014; Sies Citation2014; Andreyev et al. Citation2015; Orr et al. Citation2015; Brand Citation2016; Brand et al. Citation2016; van ‘t Erve et al. Citation2017; Crooks et al. Citation2018; Murphy and Hartley Citation2018; Watson et al. Citation2019; Sies and Jones Citation2020). The present review outlines the conditions that favor superoxide and hydrogen peroxide production by electron leak from different sites in the mitochondrial electron transport chain, and then discusses the evidence supporting roles for superoxide and hydrogen peroxide in subsequent biological effects: physiological signaling and pathological damage.

Sites of mitochondrial production of superoxide and hydrogen peroxide and their relative importance

At least eleven distinct sites in the mitochondrial electron transport chain and associated substrate dehydrogenases have been characterized and shown to generate superoxide or hydrogen peroxide at measurable rates using isolated mitochondria. The characteristics of these sites are reviewed elsewhere (Brand et al. Citation2004; Brand Citation2010; Quinlan et al. Citation2013; Brand Citation2016; Wong et al. Citation2017, Citation2019). Each site generates superoxide on the matrix side of the mitochondrial inner membrane, or within the matrix itself, and some of them may also generate hydrogen peroxide as the primary product. Superoxide dismutase-2 (SOD2; MnSOD) in the matrix converts superoxide to hydrogen peroxide, so all sites can contribute to superoxide and hydrogen peroxide levels in the mitochondrial matrix. Hydrogen peroxide (but not superoxide) can diffuse out of the mitochondria to the cytosol, so in principle all sites can also contribute to cytosolic hydrogen peroxide levels. Two sites, one in complex III and one in mitochondrial glycerol 3-phosphate dehydrogenase, produce about half of their superoxide directly into the intermembrane space (St-Pierre et al. Citation2002; Muller et al. Citation2004; Brand Citation2016) and can therefore contribute to superoxide levels in the cytosol as well as in the mitochondrial matrix. Each of these eleven sites is a potential source of superoxide or hydrogen peroxide in cells and in vivo, and might contribute to redox signaling and overt pathology. At one time it was reasonable to question whether mitochondria make any significant contribution to cellular hydrogen peroxide levels (Brown and Borutaite Citation2012), but there is now good evidence that they do: in resting cultured C2C12 myoblasts and myocytes, mitochondria are the largest single contributor to the pool of hydrogen peroxide in the cytosol (Wong et al. Citation2018; Goncalves et al. Citation2020), and in a range of cell types they contribute about 30% of the total (Fang et al. Citation2020).

Measurements of the rates and relative importance of different mitochondrial sites of superoxide and hydrogen peroxide production in cells are sparse, but four sites, sites IQ and IF in complex I, IIF in complex II, and IIIQo in complex III (where subscripts ‘Q’ and ‘F’ denote the quinone and flavin sites, respectively), dominate in mitochondria isolated from rat skeletal muscle and incubated in a medium mimicking the cytosol in muscle at rest (Goncalves et al. Citation2015, Citation2020). S1QELs (Brand et al. Citation2016) and S3QELs (Orr et al. Citation2015) are small molecules that suppress electron leak from the respiratory chain to oxygen to form superoxide and hydrogen peroxide at sites IQ and IIIQo, respectively. Importantly, they do so without inhibiting forward or reverse electron transport in the respiratory chain (Wong et al. Citation2019) and without inhibiting oxidative phosphorylation. Measurement of the acute effects of S1QELs and S3QELs on total cellular hydrogen peroxide production has enabled extension of these ex vivo observations (Goncalves et al. Citation2015, Citation2020) to intact cells and organisms, and has shown that sites IQ and IIIQo dominate the mitochondrial contribution to cytosolic hydrogen peroxide levels in cultured C2C12 myoblasts and myocytes (Wong et al. Citation2018; Goncalves et al. Citation2020). Just one site, site IQ, is responsible for at least half of the superoxide and hydrogen peroxide generated in the mitochondrial matrix in a range of cell types cells, and most of the remainder originates from site IIIQo (Wong et al. Citation2018; Fang et al. Citation2020). Site IQ is also a significant contributor to hepatic and cardiac oxidative burden in vivo (Wong et al. Citation2020). By extrapolation, the physiological and perhaps the pathological effects caused by matrix production of superoxide or hydrogen peroxide that are discussed below may be driven largely by superoxide or hydrogen peroxide derived from sites IQ and IIIQo.

Mitochondrial levels of superoxide and hydrogen peroxide

What reactions lead to elevated physiological or pathological levels of superoxide or hydrogen peroxide derived from a particular mitochondrial site? The concentration of superoxide in the mitochondrial matrix will be set by the rate of its production (reaction 3 in ) and the rate of its removal (primarily reaction 4 in ). When superoxide is produced faster than it is removed, the concentration of superoxide will rise progressively and reaction 4 will be driven progressively faster until removal occurs at the same rate as production and a steady state is achieved. In the same way, the concentration of hydrogen peroxide in the mitochondrial matrix at steady state will be set by the rate of its production (reaction 4 in , plus any direct production from mitochondrial sites) and the rate of its removal (primarily reactions 5 and 6 in ). Other factors that are commonly discussed, such as tissue specificity, protein expression level, electron transport rate, rate of oxygen consumption, mitochondrial dysfunction, respiratory chain inhibition, substrate preference, substrate concentration, calcium uptake, mitochondrial membrane potential, uncoupling, mitochondrial permeability transition pore, NAD redox state, ubiquinone redox state, allosteric modification of proteins, post-translational modification of proteins by phosphorylation or acetylation, antioxidant status, and so on, must always operate solely through changes in these proximal rates. Elevated steady-state levels of matrix superoxide or hydrogen peroxide are caused by increased production rates or decreased activity of removal pathways, or both.

Figure 1. Pathways of mitochondrial superoxide and hydrogen peroxide production and removal. X* is the reduced form of one of eleven redox centers (X) in the mitochondrial electron transport chain and associated substrate dehydrogenases; ‘*-’ denotes the species that reacts with O2, usually its single-electron reduced form, such as a semiquinone in complex I or complex III. Reactions 1-6 are discussed in the text. SOD2: superoxide dismutase-2; PRDX3: peroxiredoxin-3; GPX1: glutathione peroxidase-1; mCAT: catalase expressed transgenically in the mitochondrial matrix; *OH: hydroxyl radical (see color version of this figure at www.tandfonline.com/ibmg).

Figure 1. Pathways of mitochondrial superoxide and hydrogen peroxide production and removal. X*− is the reduced form of one of eleven redox centers (X) in the mitochondrial electron transport chain and associated substrate dehydrogenases; ‘*-’ denotes the species that reacts with O2, usually its single-electron reduced form, such as a semiquinone in complex I or complex III. Reactions 1-6 are discussed in the text. SOD2: superoxide dismutase-2; PRDX3: peroxiredoxin-3; GPX1: glutathione peroxidase-1; mCAT: catalase expressed transgenically in the mitochondrial matrix; *OH: hydroxyl radical (see color version of this figure at www.tandfonline.com/ibmg).

(i) Rates of production and removal of mitochondrial superoxide

What determines the underlying rates of production and removal? The reduction of O2 to superoxide (1-electron reduction) or hydrogen peroxide (2-electron reduction) by a relevant electron donor (X*) normally follows simple second-order chemical kinetics. Findings of little dependence on oxygen level at physiological oxygen tension (Hoffman et al. Citation2007) have been superceded (Grivennikova et al. Citation2018; Stepanova et al. Citation2019). Therefore, the rate of production of superoxide (or of hydrogen peroxide, if it is a direct product of reaction 3 in ) depends only on the concentration of X*, on the local concentration of O2, and on the rate constant k for their reaction.

The concentration of X* is determined by the size of the pool of X (for example, by the amount of the relevant respiratory complex), by the rate of X* production by upstream reductants (or nominally downstream reductants driven back up to X by reverse electron transport) (reaction 1 in ) and by the kinetics of X* consumption by downstream oxidants (reaction 2 in ). The rate of reaction 1 can be altered by the supply of electrons (which is determined, for example, by succinate or pyruvate concentration, or by matrix NADH/NAD+ ratio) and by the activities of the enzymes or electron transport chain components that pass electrons from such substrates to X. These enzymes can be regulated (for example, by the level of free Ca2+ in the mitochondrial matrix, which affects the activity of pyruvate, 2-oxoglutarate and isocitrate dehydrogenases), or their activities can be decreased by appropriate enzyme or electron transport chain inhibition. The rate of reaction 2 can be altered by changes in the downstream electron transport chain activity (for example, by the addition of respiratory poisons, by changed energy demand [which alters the redox state of downstream chain components], or by altered rate constants for electron transport). The local concentration of O2 can be altered by hypoxia or hyperoxia at the site (caused, for example, by altered oxygen supply or delivery to the tissue, or by altered respiration rate, which changes the local oxygen tension and intracellular oxygen gradients). The rate constant, k, for the reaction of X* with O2 can be altered in many ways, for example by mutations, conformational changes, post-translational modifications, or small molecules such as S1QELs and S3QELs.

Based on these principles, at any given O2 tension, the rate of superoxide and hydrogen peroxide production by site IF is determined by the amount of active complex I, the size of the matrix NADH plus NAD+ pool, and the matrix NADH/NAD+ ratio (Kussmaul and Hirst Citation2006; Treberg et al. Citation2011; Quinlan, Treberg, et al. Citation2012; Quinlan et al. Citation2014). The rate of superoxide and hydrogen peroxide production by site IQ is determined by the amount of active complex I, the matrix NADH/NAD+ ratio, the redox state of the ubiquinone pool (QH2/Q ratio), the mitochondrial membrane potential, and the mitochondrial pH gradient (Lambert and Brand Citation2004; Lambert et al. Citation2010; Treberg et al. Citation2011; Robb et al. Citation2018). The rate of superoxide and hydrogen peroxide production by site IIF is determined by the amount of active complex II, the QH2/Q ratio, and the succinate and other dicarboxylate concentrations (Quinlan, Orr, et al. Citation2012; Quinlan, Treberg, et al. Citation2012). The rate of superoxide production by site IIIQo is determined by the amount of active complex III, the QH2/Q ratio, and the mitochondrial membrane potential (Quinlan et al. Citation2011; Quinlan, Treberg, et al. Citation2012).

Very few of the electrons flowing down the respiratory chain in cells or in vivo leak prematurely to O2 to form superoxide or hydrogen peroxide. We estimate that only 0.35% do so in skeletal muscle at rest, dropping to 0.01% during exercise (Goncalves RL et al. Citation2015). These numbers give a sense of scale to the biological fluxes of electrons, but they are not useful measures of the rate of production of superoxide or hydrogen peroxide. The rate of electron transport through the electron transport chain () affects superoxide and hydrogen peroxide production only through its effects on the concentrations of various X* species (and through any effects on intracellular oxygen gradients). The level of a particular X* can be increased either by increasing reaction 1 and raising respiration rate, or by decreasing reaction 2 and lowering respiration rate. For this reason, the rate of respiration does not predict the rate of superoxide or hydrogen peroxide production. The proportion of electrons flowing down the respiratory chain that end up forming superoxide depends completely on the manipulations made during the measurements, and changes in this proportion are uninformative about the intensity of production of superoxide or hydrogen peroxide in physiology or pathology; the absolute rates or levels are much more informative.

The rate of removal of matrix superoxide (reaction 4 in ) is determined primarily by the rate of dismutation catalyzed by SOD2, the only superoxide dismutase normally present in the matrix (Weisiger and Fridovich Citation1973). Spontaneous dismutation can also be significant (and is the dominant pathway when matrix superoxide levels rise in SOD2 null cells). Rates of reaction with other species are normally quantitatively minor (although potentially of great biological significance), except for the reaction with nitric oxide (NO) to form peroxynitrite, which can be a significant flux when NO production rate is high.

(ii) Rates of production and removal of mitochondrial hydrogen peroxide

Production of hydrogen peroxide in the matrix is normally catalyzed mostly by SOD2, with some contribution from spontaneous dismutation (together comprising reaction 4 in ) plus any direct production from mitochondrial sites. Any diffusion into the mitochondrial matrix from the rest of the cell will also contribute; when this happens extensively mitochondria can be considered to be sinks of hydrogen peroxide, potentially acting as a buffer of cytosolic hydrogen peroxide levels (Mailloux Citation2018). There are several mechanisms for removal of hydrogen peroxide from the matrix. The main pathway is thought to be catalyzed by peroxiredoxin-3 (PRDX3), which is targeted to the mitochondrial matrix (Chae et al. Citation1999) and has been calculated to clear 90% of matrix hydrogen peroxide (Cox et al. Citation2009), using reducing equivalents from thioredoxin-2 generated by thioredoxin reductase-2 using matrix NADPH. Under conditions of stress, peroxiredoxin 6 may also be trafficked to the mitochondrial matrix (Eismann et al. Citation2009). Glutathione peroxidases, primarily GPX1, also play a role, using reducing equivalents from matrix glutathione generated by glutathione reductase using matrix NADPH. Together these make up reaction 5 in . Diffusion of matrix hydrogen peroxide to the cytosol (reaction 6 in ) is also quantitatively and biologically significant. Reactions with other species are normally quantitatively minor (although potentially of great biological significance). Since superoxide in the matrix is converted almost quantitatively into hydrogen peroxide by SOD2 or spontaneous dismutation, the addition of pure SOD mimetics will not alter the rate of hydrogen peroxide production or the matrix concentration of hydrogen peroxide (unless NO is very high, allowing a competition for superoxide to ensue) but will lower the steady-state concentration of superoxide (see below). The kinetics of hydrogen peroxide removal from the matrix can be altered in several ways, for example by addition of catalase mimetics, by changed glutathione levels in the matrix (Mari et al. Citation2009), or by mutations or inhibitors that affect other matrix antioxidant systems. Some of the best evidence for a significant role for mitochondrial hydrogen peroxide in driving different phenotypes comes from experiments using genetically engineered expression of catalase in the mitochondrial matrix (mCAT), where it is not normally found (Dai et al. Citation2014, Citation2017).

Mechanisms driving biological effects of mitochondrial superoxide or hydrogen peroxide

The mechanism by which a physiological change or a pathological insult increases the mitochondrial concentrations of superoxide and hydrogen peroxide derived from a specific site or sites and leads to biological effects will be by altering one or more of the variables discussed above. In ischemia-reperfusion injury the mechanism is thought to be altered substrate supply, as succinate built up during ischemia is used on reperfusion to drive reverse electron transport to reduce sites IQ and IF, leading to increased production of superoxide and hydrogen peroxide, increased levels of these species, and reperfusion injury (Chouchani, Pell, et al. Citation2016). In various mitochondrial diseases it may be simply inhibition of the reoxidation of X* caused by a loss-of-function mutation in downstream electron transport complexes or the ATP synthase, or it may be a gain-of-function change in the rate constant for superoxide production, as in certain succinate dehydrogenase mutations in Caenorhabditis elegans (Adachi et al. Citation1998). In acetaminophen toxicity it is largely the decrease in removal pathways initiated by depletion of glutathione by direct adduct formation, and inhibition of mitochondrial glutathione peroxidases, both caused by reactive metabolites of acetaminophen (Du et al. Citation2016).

Altering the kinetics of superoxide and hydrogen peroxide removal using transgenes and antioxidants

A widespread problem in the literature is the lack of precision in setting up hypotheses, in particular, the discussion of various reactive oxygen species as if they were all equivalent and interchangeable – they are not (Murphy et al. Citation2011; Sies and Jones Citation2020). When the nature and reactions of a particular reactive oxygen species are considered, it becomes obvious that different manipulations, such as SOD2 overexpression, mCAT expression, addition of the targeted lipid peroxidation chain-breaker mitoQ, or addition of the glutathione precursor N-acetylcysteine are not equivalent and cannot meaningfully be lumped together as “anti-ROS” or “antioxidant” treatments (Davis et al. Citation2001; Murphy et al. Citation2011; Lark et al. Citation2015; Sies and Jones Citation2020). It is important to emphasize that decreases (including complete knockout), or increases (even to saturating levels) in the activities of the enzymes that remove superoxide and hydrogen peroxide do not usually have much effect on the rates of production of superoxide or hydrogen peroxide by mitochondria. Instead, such changes alter the steady-state concentrations of these species in the mitochondrial matrix.

Consider again. The superoxide concentration in the mitochondrial matrix is set by a steady state in which its rates of production and consumption are equal. Following some arbitrary imposed increase in mitochondrial superoxide production (from complex I or other sites facing the matrix; reaction 3), the matrix superoxide concentration will rise until, in a new steady-state, it drives a new rate of consumption that is equal to the new rate of production. Because they operate very far from equilibrium, the reactions that generate superoxide are essentially irreversible and insensitive to superoxide concentration, so there will be no direct secondary or compensatory change in their rates in response to the rising concentration of their product, superoxide. Some superoxide will react with other molecules (proteins, lipids, NO, mtDNA etc.; the minor fluxes) but, overwhelmingly, the major consumers are the conversion of superoxide to hydrogen peroxide by SOD2-catalyzed and spontaneous dismutation (reaction 4). Under cellular conditions these consumer reactions are again essentially irreversible and independent of the concentration of their product, hydrogen peroxide.

These properties have two important consequences. First, because the dismutation rate (reaction 4) is so much higher than the rates of the other reactions (minor fluxes), the superoxide that is generated is converted almost quantitatively to hydrogen peroxide, at a rate determined by the superoxide generation rate, and (because of spontaneous dismutation) independent of the activity of SOD2. Second, the steady-state concentration of superoxide (and hence the rates of the minor side-reactions) is very sensitive to the activity of SOD2. If SOD2 is activated (or pure SOD mimetics are added), matrix superoxide concentration will drop until the steady state is reestablished at almost the same consumption rate and hydrogen peroxide production rate as before the change, but at a lower level of superoxide and hence lower rates of the quantitatively minor side-reactions. If SOD2 activity is decreased, such as in a SOD2 heterozygous animal, matrix superoxide concentration will be higher but hydrogen peroxide production will hardly be affected (it will slow very slightly as the side-reactions from superoxide increase and carry a little more of the total flux). In the limit, if SOD2 is removed completely, as in a SOD2 knockout, matrix superoxide will rise until it reaches a high level sufficient to drive all the flux to hydrogen peroxide by spontaneous dismutation (and the relatively minor rates of the side-reactions with other molecules will be maximal). Thus, modification or mimicry of SOD2 activity will change matrix superoxide concentration, with potentially significant effects of changes in rate of the superoxide side-reactions, but will have almost no effect on matrix or cellular hydrogen peroxide concentration. See (Gardner et al. Citation2002) for a simple quantitative model of how altered SOD2 activity can increase, not affect, or decrease hydrogen peroxide production; only in the few exceptional cases in which endogenous pathways that consume matrix superoxide without forming hydrogen peroxide run at a significant rate will increased SOD2 activity cause increased hydrogen peroxide production (by successfully competing for the pool of superoxide). Thus, the major fluxes from superoxide to hydrogen peroxide will be essentially independent of manipulations in the kinetics of reaction 4. Instead, the steady state level of superoxide and therefore the minor fluxes of its other reactions will be very sensitive to manipulations in the kinetic properties of reaction 4, for example by altered expression of SOD2 or the addition of SOD mimetics.

Similar considerations apply to the steady state that determines the matrix hydrogen peroxide concentration. The arbitrary imposed increase in mitochondrial superoxide production will drive a matching rate of hydrogen peroxide production, and the matrix hydrogen peroxide concentration will rise until it drives a new rate of consumption, mostly by peroxiredoxin-3 as discussed above, that is equal to the new rate of production. Other than diffusion to and from the cytosol, the producers and consumers of hydrogen peroxide (SOD2 and spontaneous dismutation; peroxiredoxin-3, mCAT, glutathione peroxidases, catalase mimetics) run far from equilibrium and are therefore essentially irreversible and product-independent, so the same rules apply: modulations of the consumers (reaction 5) will affect the hydrogen peroxide concentration in the matrix and the rate of escape of hydrogen peroxide from the mitochondria (reaction 6), but will have virtually no effect on hydrogen peroxide production rate (reaction 4) or on matrix superoxide levels. Thus, modification of peroxiredoxin-3 activity or expression of mCAT will change matrix (and therefore cytosolic) hydrogen peroxide concentration, with potentially significant effects of changes in rate of the hydrogen peroxide side-reactions, but will have no effect on matrix superoxide production rate or concentration.

Different antioxidants that act in the mitochondria will have different effects depending on their properties. Those that react directly with superoxide will lower superoxide level by adding a new superoxide consumer to the steady state. This will decrease the rates of the superoxide side-reactions, and lower hydrogen peroxide level by diverting superoxide flux. Those that react directly only with hydrogen peroxide will not affect superoxide concentration, but will decrease hydrogen peroxide concentration (by adding a new hydrogen peroxide consumer to the steady state) and decrease the rates of the hydrogen peroxide side-reactions. Antioxidants that affect downstream reactions, such as those that act as chain-breakers in lipid peroxidation reactions, will affect lipid peroxidation and downstream reactions, but will have no direct effect on superoxide and hydrogen peroxide steady state levels or the biological effects they drive. S1QELs and S3QELs will decrease premature electron flow to O2 and lower the levels of both superoxide and hydrogen peroxide, as will oxidation of respiratory chain components. If a particular phenotype is caused by matrix superoxide and not by matrix hydrogen peroxide, etc. (see ), then its responses to different “antioxidants” may be very different, sowing confusion unless the nature of “ROS” is more thoughtfully specified.

It follows from these considerations that different manipulations of “ROS” levels can be much more diagnostic of which species is causal than is commonly assumed. If a phenotype is prevented by SOD2 overexpression and S1QEL or S3QEL addition but not by peroxiredoxin-3 overexpression or mCAT expression, then it is caused by matrix superoxide and not by matrix hydrogen peroxide. Similar arguments apply to other combinations of responses to genetic and pharmacological interventions. The possible responses of a particular phenotype to relevant experimental manipulations and the interpretations are shown in . Thus, different “antioxidants” can have profoundly different effects, and general conclusions that take no account of this property can be very misleading, and miss important implications that could be extracted from a particular dataset.

Figure 2. Identification of mitochondrial sources of superoxide and hydrogen peroxide that cause a given phenotype using different manipulations that affect that phenotype.

Figure 2. Identification of mitochondrial sources of superoxide and hydrogen peroxide that cause a given phenotype using different manipulations that affect that phenotype.

Proximal and more distal targets of matrix superoxide and hydrogen peroxide

What are the immediate targets of superoxide and hydrogen peroxide in the mitochondrial matrix? The chemical reactivities of superoxide and hydrogen peroxide are very different, so they have different molecular targets (Collin Citation2019; Sies and Jones Citation2020). The negatively charged superoxide radical anion, O2•−, is reactive and a good oxidant of positively charged Fe-S centers in proteins. The classic example is the [4Fe-4S]2+ center of the tricarboxylic acid cycle enzyme aconitase in the mitochondrial matrix, which is very sensitive to inactivation by superoxide (and 10,000 times less sensitive to hydrogen peroxide) (Gardner Citation2002; Imlay Citation2006; Castro et al. Citation2019). Its inactivation may lead directly to inappropriate lipogenesis (Crooks et al. Citation2018). Succinate dehydrogenase (complex II) in the matrix is also extremely sensitive to superoxide (Melov et al. Citation1999; Powell and Jackson Citation2003; Lustgarten et al. Citation2011; Brand et al. Citation2016), probably by inactivation of one of its Fe-S centers. The protonated form of superoxide, perhydroxyl radical (HOO), is uncharged and lipid-soluble. It reacts with polyunsaturated fatty acyl side-chains of mitochondrial phospholipids, particularly cardiolipin (because it tethers four fatty acyl side-chains in close proximity), to initiate cascades of lipid hydroperoxide formation that alter the functions of the mitochondrial inner membrane, and also react with mitochondrial DNA and membrane proteins.

In contrast to superoxide, hydrogen peroxide is relatively stable, but can rapidly oxidize specific reactive protein thiolates (which are more prominent in the mitochondrial matrix because of its alkaline pH), to the sulfenic and sulfinic acids, and lead to the formation of disulfides. An important thiolate target of hydrogen peroxide in the mitochondrial matrix may be cysteine residues on pyruvate dehydrogenase kinase 2 (PDHK2). Their oxidation by hydrogen peroxide inhibits PDHK2 activity, resulting in activation of the pyruvate dehydrogenase complex and increased oxidation of pyruvate (Hurd et al. Citation2012). Hydrogen peroxide is membrane-permeant, so it will diffuse to the cytosol where, mostly through modification of protein thiols, it can orchestrate changes in protein activity through activation of MAP protein kinases (Siauciunaite et al. Citation2019) and the HIF1α and other pathways (Klimova and Chandel Citation2008; Diebold and Chandel Citation2016), and changes in gene transcription through redox-regulated transcription factors such as c-Jun, ATF2, ATF4, NFkB, Nrf2 and p53 (Siauciunaite et al. Citation2019).

These reactions of superoxide and hydrogen peroxide generate secondary species that are even more reactive than the parent species. Oxidation of Fe-S centers by superoxide releases Fe2+ into the matrix, which in turn catalyzes Haber-Weiss chemistry to generate OH radical from superoxide and hydrogen peroxide, and Fenton chemistry to generate OH radical from hydrogen peroxide. OH radicals are highly reactive but have only weak specificity, and react at diffusion-limited rates with proteins to inactivate them, mtDNA bases to cause mutation, and polyunsaturated fatty acyl side-chains of mitochondrial phospholipids. Matrix superoxide can also activate mitochondrial uncoupling proteins (Echtay, Murphy, et al. Citation2002; Echtay, Roussel, et al. Citation2002), perhaps through lipid peroxidation reactions (Brand et al. Citation2004). HOO reacts rapidly with nitric oxide to limit its concentration and signaling potential, forming peroxynitrite (ONOO-), a powerful tyrosine-nitrating and methionine-oxidizing agent. Hydrogen peroxide reacts with bicarbonate to form peroxymonocarbonate (HCO4), which reacts with thiols much more readily than the parent hydrogen peroxide (Collin Citation2019; Sies and Jones Citation2020).

Biological effects

As laid out in the Introduction, there are large and unwieldy bodies of literature on the roles of mitochondria-derived superoxide and hydrogen peroxide in cellular signaling and pathology. lists phenotypes that are probably driven by such signaling pathways. Pathologies that may be driven by mitochondrial production of superoxide or hydrogen peroxide are listed in .

Table 1. Published evidence that particular physiological processes are driven or modified by mitochondrial production of superoxide and/or hydrogen peroxide, according to the numbered criteria in .

Table 2. Published evidence that particular pathologies are driven or modified by mitochondrial production of superoxide and/or hydrogen peroxide, according to the numbered criteria in .

Much of the evidence for involvement of mitochondrial superoxide and/or hydrogen peroxide in a particular signaling or pathological phenotype is circumstantial, making the conclusions tentative. First, most of the studies cited in the reviews listed above are associative rather than mechanistic; they show that some signal attributed to mitochondrial ROS (such as the expression of an antioxidant protein, or the response of a fluorescent probe) is increased when the physiological signaling pathway or pathology is activated, or decreased when it is inactive. Such associations are suggestive, but they do not distinguish between causal and bystander or downstream roles of mitochondrial ROS. Second, when these studies probe mechanism by manipulating mitochondrial ROS production they often use mitochondrial poisons or genetic manipulations that alter not only mitochondrial superoxide and hydrogen peroxide production but also cellular energetics and metabolite concentrations. This makes such studies theoretically unreliable, because it is hard, if not impossible, to unscramble effects of mitochondrial superoxide and hydrogen peroxide from the simultaneous effects of altered energetics and metabolism caused by the manipulation. Third, many of the cited studies use cellular probes for reactive oxygen species that suffer from many serious drawbacks, such as lack of specificity for particular ROS (or even for ROS over other redox-active reactants), interference with the redox status of the target cell (for example by acting as pro-oxidants), and inadequate controls for effects of changes in plasma membrane potential and mitochondrial membrane potential on the accumulation of the mitochondria-targeted probes, etc. (Wardman Citation2007; Murphy et al. Citation2011; Kalyanaraman et al. Citation2012, Citation2017; Xiao and Meierhofer Citation2019), making them experimentally unreliable.

Strength of different types of evidence that mitochondria-derived superoxide or hydrogen peroxide drives biological effects

Given the difficulties with association studies, what types of evidence can give robust support to the hypothesis that mitochondrial production of superoxide or hydrogen peroxide is an important driver of a particular physiological signaling pathway or a particular pathology, or, conversely, can provide strong evidence against that hypothesis? Ideally, supportive evidence should show that production is both necessary (i.e. preventing a rise in superoxide or hydrogen peroxide concentration in the mitochondrial matrix prevents the engagement of the signaling pathway or the induction of the pathology in response to its normal initiators) and sufficient (i.e. artificially triggering a rise in the level of matrix superoxide or hydrogen peroxide triggers the signal generation or pathology). In general, pharmacological or genetic interventions that decrease or increase concentrations of superoxide or hydrogen peroxide specifically in the mitochondrial matrix without affecting other reactions such as oxidative phosphorylation are required. Each of these interventions has advantages and drawbacks.

The advantage of pharmacological interventions is that they can be applied to essentially any experimental system without the need for complex genetic preparation, and they work acutely (within seconds to hours), without necessarily giving the system time to make compensatory changes in gene expression. The disadvantage is that they require a well-characterized pharmacologically-active molecule with specificity for the target and with suitable properties such as solubility, chemical and biological stability, and membrane permeability. The chief potential drawback of pharmaceutical approaches is always non-specificity, in this case the possibilities of residual inhibition of the electron transport chain or completely off-target effects at high concentrations always have to be borne in mind. Although short-term responses to well-characterized pharmaceuticals can be interpreted with little ambiguity, long-term treatment with a pharmacological reagent allows time for compensatory signaling and gene expression changes (such as down-regulation of endogenous antioxidant pathways) that can complicate interpretations.

The advantage of genetic manipulations is specificity; with appropriate design, a particular protein can be overexpressed or knocked down in a specified tissue and subcellular compartment, at a time of the experimenter’s choosing. The chief drawback of genetic approaches is always the poor time resolution. Relatively long times (several hours, days or weeks) are needed to allow overexpression, or turnover of existing protein in knock-down experiments, which allows a myriad of compensatory changes in the expression of other genes, or changes in development, long-term damage etc. These consequent changes always generate ambiguity in the interpretation of phenotypes caused by making changes in the expression of a gene of interest. In addition, overexpressed proteins may aggregate, or bind and alter partner molecules, causing unexpected secondary effects, and the target protein might have other quite separate unforeseen functions, complicating the analyses.

summarizes the different manipulations that can provide robust support, ranked according to how proximal they are to the primary variables of interest: matrix superoxide and hydrogen peroxide concentrations. These manipulations are used below as criteria to judge the strength of published evidence in supporting the hypothesis that mitochondrial production of superoxide and/or hydrogen peroxide drives physiological signaling pathways () or pathologies ().

Figure 3. Criteria for assessing whether mitochondrial production of superoxide and/or hydrogen peroxide drives a biological effect, ranked by reliability. For details and justifications see the text.

Figure 3. Criteria for assessing whether mitochondrial production of superoxide and/or hydrogen peroxide drives a biological effect, ranked by reliability. For details and justifications see the text.
  • 1. The inhibition of a phenotype by addition of well-characterized S1QELs or S3QELs is the best evidence that the phenotype is caused by mitochondrially-generated superoxide or hydrogen peroxide. The general problem with directly manipulating the production of superoxide (or hydrogen peroxide) (reaction 3 in ) is that such manipulation using conventional reagents modifies the flow of electrons in the respiratory chain, causing changes in energetics that can dominate the observed effects. For this reason effects of mitochondrial respiratory poisons such as rotenone, myxothiazol, or antimycin, effects of uncouplers of oxidative phosphorylation, or effects of genetic modulations to the activity of respiratory complexes, such as mutations in complex I or the Rieske FeS center of complex III, are extremely unreliable guides to the biological roles of mitochondrial superoxide and hydrogen peroxide. This problem has been overcome by the discovery, using chemical screens, of small molecules that suppress electron leak to oxygen without inhibiting electron transport in the respiratory chain. S1QELs (Brand et al. Citation2016) suppress site IQ electron leak, and S3QELs (Orr et al. Citation2015) suppress site IIIQo electron leak (Orr et al. Citation2013, Citation2015; Brand Citation2016; Brand et al. Citation2016; Wong et al. Citation2017). Specific suppressors of the other sites of superoxide and hydrogen peroxide production have not yet been published. By directly decreasing formation of superoxide (and perhaps hydrogen peroxide), S1QELs and S3QELs decrease matrix levels of both superoxide and its product, hydrogen peroxide. Unlike conventional antioxidants they have the advantages of site-selectivity and of potentially fully-preventing superoxide production from a particular site rather than mopping up after superoxide or hydrogen peroxide have had effects (Watson et al. Citation2019). The best of the published S1QELs and S3QELs have high potencies (low IC50s) in mitochondrial assays: S1QEL1.1, 70 nM (Brand et al. Citation2016); S3QEL3, 350 nM (Orr et al. Citation2015), and cause no inhibition of cell growth or the electron transport chain at 20 times higher concentrations. Other S1QELs with lower potencies have also been discovered: anethole trithione (5-[p-methoxyphenyl]-1,2-dithiole-3-thione, Sulfarlem®, OP2113), IC50 = 10–26 µM (Boucard et al. Citation2019; Detaille et al. Citation2019), and Imeglimin, IC50 = ∼100 µM (Detaille et al. Citation2016). However, it is unclear whether the S1QEL activities of anethole trithione (a known slow-release H2S donor (Wallace et al. Citation2018)) or Imeglimin are the only cause of their biological effects.

  • 2. Excellent evidence for involvement of mitochondrially-generated superoxide in a phenotype comes from manipulating its removal (reaction 4 in ). As outlined above, SOD2 is the only superoxide dismutase normally present in the matrix (Weisiger and Fridovich Citation1973) and it provides the only significant catalyzed pathway for removal of matrix superoxide; the other pathway that can reach high rates is spontaneous dismutation. Overexpression of SOD2 will lower the steady-state superoxide concentration in the matrix, usually without affecting matrix hydrogen peroxide production or concentration, and will decrease effects caused by matrix superoxide. One caveat to SOD2 overexpression studies is that there needs to be an adequate supply of Mn and lack of excessive Fe, otherwise the overexpressed apoenzyme may incorporate Fe instead of Mn, making it a peroxidase with damaging pro-oxidant effects (Ganini et al. Citation2018). Conversely, partial or complete SOD2 knockout will raise matrix superoxide concentration, amplifying its effects. Excellent evidence for a significant physiological or pathological role of matrix superoxide comes from prevention of a phenotype by SOD2 overexpression, or exacerbation by SOD2 knockdown. Constitutive knockout of SOD2 is neonatally lethal in mice (Li, Huang, et al. Citation1995), illustrating that excessively raised matrix superoxide concentrations can exert dramatic pathological effects.

  • 3. Excellent evidence for involvement of matrix-derived hydrogen peroxide in a phenotype comes from manipulating its removal (reaction 5 in ). As outlined above, PRDX3 may normally remove 90% of matrix hydrogen peroxide (Cox et al. Citation2009). Overexpression of PRDX3 will lower the steady-state hydrogen peroxide concentration in the matrix, without affecting matrix superoxide production or concentration, and will decrease effects caused by matrix hydrogen peroxide. Conversely, partial or complete PRDX3 knockout will raise matrix hydrogen peroxide concentration, amplifying its effects. Other systems also remove matrix hydrogen peroxide, in particular diffusion to the cytosol and to other organelles (reaction 6 in ), where it can be degraded by catalases and peroxidases. GPX1 is also active in the matrix, but the majority (about 90%) is located in the cytosol and it also catalyzes the reduction of various other hydroperoxides (Asayama et al. Citation1994; Esworthy et al. Citation1997; Ho et al. Citation1997), making it a poor candidate for driving specific alterations in matrix hydrogen peroxide concentration. Strong evidence for a significant physiological or pathological role of matrix hydrogen peroxide comes from prevention of a phenotype by PRDX3 overexpression, or exacerbation by PRDX3 knockdown. However, PRDX3 is not fully specific for hydrogen peroxide, but also removes other hydroperoxides (Peskin et al. Citation2010), making interpretations less secure than those for superoxide based on manipulations of SOD2. Some of the best evidence for a significant role for mitochondrial hydrogen peroxide in driving different phenotypes comes from experiments using genetically engineered expression of catalase in the mitochondrial matrix (mCAT) (Dai et al. Citation2014, Citation2017). Expression of mCAT will decrease hydrogen peroxide but not superoxide concentrations in the matrix, and has been used extensively to provide high-quality evidence for involvement of matrix-derived hydrogen peroxide in numerous pathologies (see ).

  • 4. Good evidence for the involvement of mitochondrially-generated superoxide in a phenotype comes from protective effects of mitochondria-targeted SOD mimetics (and exacerbation by mitochondria-targeted superoxide-generators). mitoTEMPO and mitoTEMPOL are piperidine nitroxides attached to a lipophilic cation, enabling them to pass through lipid bilayers and accumulate several hundred-fold in mitochondria. These and related or equivalent molecules are mitochondria-targeted SOD mimetics. However they have sub-optimal chemical specificity because they are also peroxidases and lipid peroxidation chain terminators (Soule et al. Citation2007) and they interact with the electron transport chain at ubiquinone (Trnka et al. Citation2008), making them less definitive than we would wish. Nonetheless, experiments with mitoTEMPO provide some of the most extensive pharmacological evidence in the literature for the roles of matrix superoxide in physiology and pathology (see and ). Pro-oxidants such as mitochondria-targeted paraquat, mitoPQ (Robb et al. Citation2015) should also be useful, but their specificity is unclear and they have not yet been used extensively.

  • 5. Mitochondria-targeted antioxidants that work more distally from superoxide and hydrogen peroxide production can also provide useful evidence for the involvement of generic matrix oxidants in different phenotypes. MitoQ and SkQ and their derivatives are quinones attached to a lipophilic cation, enabling them to pass through lipid bilayers and accumulate several hundred-fold in mitochondria (Skulachev Citation2013; Murphy Citation2016). They probably act as primarily as lipid peroxidation chain terminators (Asin-Cayuela et al. Citation2004; Murphy Citation2016), preventing many of the downstream effects of matrix superoxide and hydrogen peroxide. Because they are redox cyclers, there is always the possibility of pro-oxidant effects, and at higher concentrations the triphenylphosphonium targeting group can inhibit electron transport. SS-31 (elamipretide, Bendavia) is a tetrapeptide that appears to act as an targeted antioxidant through its interaction with cardiolipin, which causes SS-31 to accumulate in mitochondrial membranes (Szeto Citation2006, Citation2014; Szeto and Birk Citation2014). Its mode of action is unclear, but it may modulate mitochondrial membrane surface charge (Mitchell et al. Citation2020) and decrease cytochrome c-catalyzed cardiolipin peroxidation (Birk et al. Citation2013). Such mitochondria-targeted antioxidants have been used extensively to show protective effects in many pathologies (Smith and Murphy Citation2010; Dai et al. Citation2014; Zielonka et al. Citation2017; Szeto and Liu Citation2018), providing a large body of relevant evidence (see ).

  • 6. In principle, gene-association studies are a powerful way to identify pathologies that are modified by genetic variants in relevant proteins, specifically SOD2 and PRDX3 (and to a much lesser extent because of its dual matrix and cytosolic location and lower quantitative contribution to hydrogen peroxide removal from the matrix, GPX1). Gene-association studies have the great advantage that, unlike the cell and animal studies that underlie the criteria above, they can directly link mitochondrial matrix superoxide and hydrogen peroxide levels to specific human diseases. Many such gene-association studies are cited in . However, it is essential that such studies are adequately powered to identify small effects in high background noise. This is seldom the case, and many of the meta-analyses fail to confirm the initial reports, or show only small effects. For this reason, gene-association studies are not highly rated as a criterion in .

There are also weaker suggestive criteria that have not been used in and :

  • 7. Inhibition of a phenotype by knock-in or activation by knock-down of other mitochondrial antioxidant proteins can suggest that the phenotype is driven by ROS (from unspecified sites). The NADP-specific isocitrate dehydrogenase (IDH2) is not essential for citric acid cycle flux (Williamson and Cooper Citation1980), so unlike other citric acid cycle enzymes it can be knocked down or out to affect matrix NADPH and glutathione redox state (Kim, Baek, et al. Citation2019) without necessarily causing confounding changes in citric acid cycle flux and bioenergetics. Lowering the level of reduced glutathione in the mitochondrial matrix will raise hydrogen peroxide levels and is associated with numerous pathologies (Mari et al. Citation2009). Similarly, alterations in the thioredoxin system by manipulating the matrix enzymes thioredoxin-2 (TRX2) (Spyrou et al. Citation1997) or thioredoxin reductase-2 (TRXR2) (Lee et al. Citation1999) will have effects. However, the effects of these manipulations on superoxide and hydrogen peroxide levels in the matrix are rather indirect. GPX1, PRDX5 and GPX4 are present in several compartments, making manipulations of their activities less definitive in the present context. In addition, PRDX5 detoxifies mostly alkyl hydroperoxides and GPX4 detoxifies mostly membrane hydroperoxides (Collin Citation2019; Sies and Jones Citation2020).

  • 8. General antioxidants provide weaker circumstantial evidence. Inhibition of a phenotype by untargeted antioxidants such as N-acetylcysteine (NAC), and SOD/catalase mimetics such as the salen Mn complexes (EUKs) (Doctrow et al. Citation2012; Batinic-Haberle et al. Citation2018; Bonetta Citation2018), suggests that the phenotype may be driven by cellular ROS, although the exact targets and mechanisms are very unclear.

  • 9. The levels of the mRNA or the enzymic activities of various matrix antioxidant enzymes, such as SOD2, PRDX3, or GPXs, are often reported, as are the intensities of various markers of ROS levels, such as fluorescent probes, protein oxidation, or engagement of downstream ROS-driven signaling pathways. Association of these signals with a phenotype is helpful in drawing attention to possibilities or in strengthening the circumstantial case for ROS involvement, but by themselves they do not distinguish causes from effects, so are excluded from and . Effects of pharmacological inhibition or genetic manipulation of respiratory chain proteins, or addition of uncouplers, are deeply confounded by potential effects on energetics and cannot be relied upon, so are also excluded.

Redox signaling

There is widespread agreement that cells use mitochondrial ROS as a redox signal in many different physiological contexts (see the Introduction). Using the criteria set out in as a guide, what is the evidence that superoxide and/or hydrogen peroxide generated in the mitochondrial matrix acts as a redox signal to drive different physiological responses? The relevant literature is collated in and ranked and made easier to overview in . Note that the division between physiological responses () and pathologies () is somewhat arbitrary. For example, cell proliferation and cell migration are required for normal cell growth and development, but are also features of pathologies such as cancers. Inflammation is required for responses to infection, but chronic inflammation plays a part in many pathologies.

Figure 4. Physiological processes in which mitochondrial superoxide and/or hydrogen peroxide are implicated. Ranked representation of . Entries are ranked in order of the criteria shown in (and then alphabetically for entries with the same score). Green cells indicate a preponderance of papers reporting an effect, orange cells (marked “.....”) indicate a preponderance of papers reporting no effect, and white cells indicate that no relevant papers were identified (see color version of this figure at www.tandfonline.com/ibmg).

Figure 4. Physiological processes in which mitochondrial superoxide and/or hydrogen peroxide are implicated. Ranked representation of Table 1. Entries are ranked in order of the criteria shown in Figure 3 (and then alphabetically for entries with the same score). Green cells indicate a preponderance of papers reporting an effect, orange cells (marked “.....”) indicate a preponderance of papers reporting no effect, and white cells indicate that no relevant papers were identified (see color version of this figure at www.tandfonline.com/ibmg).

and show that there is considerable evidence that mitochondrial generation of superoxide and hydrogen peroxide has an important role in a range of physiological responses of cells. Many of these responses are related to cellular stress management, and overlap with each other. They include propagation of ER stress, apoptosis, autophagy, cellular senescence and wound healing, HIF1α signaling, and the adaptive and innate immune responses. A second group of physiological responses is related to the cell cycle, cell proliferation and growth, cell migration, and differentiation. There is sporadic evidence for several other responses, some of which may be related to the first two groups (hormetic defense is related to stress responses, synaptic pruning is related to apoptosis, and effects on the gut microbiome may be related to immunity).

Pathology

There is widespread agreement that mitochondria are able to generate damaging amounts of ROS in many different pathological contexts (see the Introduction). What is the strength and weight of the evidence that superoxide and/or hydrogen peroxide generated in the mitochondrial matrix drives a particular pathological response? The relevant literature (following the criteria set out in ) is diverse and comes partly from cell models, partly from animal models, and partly from clinical studies. It is collated in and ranked and made easier to overview in .

Figure 5. Pathologies in which mitochondrial superoxide and/or hydrogen peroxide are implicated. Ranked representation of . Entries are ranked in order of the criteria shown in (and then alphabetically for entries with the same score). Green cells indicate a preponderance of papers reporting an effect, pale green cells indicate mixed evidence, orange cells (marked “….”) indicate a preponderance of papers reporting no effect, and white cells indicate that no relevant papers were identified (see color version of this figure at www.tandfonline.com/ibmg).

and list more than 100 pathologies that may be driven by mitochondrial superoxide or hydrogen peroxide according to the literature organized using the criteria in . Some of the pathologies overlap and the divisions are somewhat arbitrary, but for 30-40 of these pathologies there are multiple lines of strong experimental evidence for a role of mitochondrial superoxide or hydrogen peroxide in cell and animal models, and sometimes in clinical studies. The top half of the entries in can be crudely grouped into a smaller number of general categories. Most of the entries in the bottom half of also fit into these general categories. The categories are:

  • Metabolic diseases (diabetes, obesity and their complications)

  • Cardiovascular diseases (hypertension, atherosclerosis, cardiomyopathy, heart failure)

  • Inflammatory diseases

  • Cancer (incidence, growth, metastasis)

  • Neurological diseases (cognitive, motor, degenerative, vision, hearing, neuropathies)

  • Ischemia/reperfusion injuries

  • Aging and aging-associated diseases

  • External insults (drugs, radiation, trauma, infection)

  • Genetic diseases

As with signaling responses, many of these pathologies overlap and the divisions between them are sometimes arbitrary. For example, inflammation is an underlying theme in most of the other categories such as metabolic diseases, neurological diseases, and aging. The greatest risk factor for most of the categories is aging, underscoring the connections between them. In addition, the presence of one disease often precipitates or exacerbates another, e.g. diabetic cardiomyopathy. External insults often cause damage that triggers other pathologiess (radiation and drugs can cause aging and inflammation, heart disease, cognitive decline, etc.). Nonetheless, it is clear that mitochondrial superoxide and hydrogen peroxide have a strong causal connection to a wide range of interconnected pathologies.

There are pathologies for which there is literature showing effects of manipulation of SOD2, but little or no literature showing effects of PRDX3 or expression of mCAT (stroke, ischemia/reperfusion injury, some cancers, acetaminophen and methamphetamine toxicity, some eye diseases, periodontitis, etc.). This asymmetry suggests the possibility that these conditions are driven by matrix superoxide but not by matrix hydrogen peroxide (see above, ), although it may simply reflect a lack of studies reporting effects of manipulations that alter matrix hydrogen peroxide concentration. Conversely, there are pathologies for which there is literature showing effects of manipulation of PRDX3 or expression of mCAT but no literature showing effects of manipulation of SOD2 (pre-eclampsia, heart failure with preserved ejection fraction, ataxia-telangiectasia, arterial and lung diseases, reproductive aging, etc.) suggesting the possibility that these conditions are driven by matrix hydrogen peroxide but not by matrix superoxide (with the same caveat about absent literature).

Conclusions

Within the huge literature on ROS, signaling, and disease, much of the evidence is pertinent and strong, showing clearly that elevated mitochondrial matrix superoxide and/or hydrogen peroxide concentrations are both necessary and sufficient to drive a wide range of physiological responses and pathological outcomes. Necessary, at least partially, because lowering matrix superoxide or hydrogen peroxide concentration by suppressing their formation or enhancing their disposal (or their downstream effects) is protective. Sufficient, at least partially, because raising matrix superoxide or hydrogen peroxide concentrations by compromising their disposal can recapitulate the phenotype. This conclusion reflects the opening paragraph of the present review: riding the tiger of abundant energy release carries the risk of premature electron escape to oxygen to form ROS that can drive damage and disease – the tiger may bite you. Increased understanding of the roles of matrix superoxide and hydrogen peroxide in these pathologies can direct and illuminate the development of therapeutic approaches to treat them while avoiding untoward interference with the beneficial signaling pathways. There is a great opportunity for improved S1QELs, S3QELs and suppressors of electron leak to oxygen at other sites, improved mitochondria-targeted SOD and catalase mimetics, and improved modifiers of downstream effects such as lipid peroxidation and FeS and thiol modification to be developed to treat the pathologies listed in . Perhaps we can eventually cage our tiger.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

Work in my laboratory is supported by Calico Life Sciences LLC (South San Francisco, CA) and the Buck Institute for Research on Aging (Novato, CA).

References

  • Abbasi M, Daneshpour MS, Hedayati M, Mottaghi A, Pourvali K, Azizi F. 2018. The relationship between MnSOD Val16Ala gene polymorphism and the level of serum total antioxidant capacity with the risk of chronic kidney disease in type 2 diabetic patients: a nested case-control study in the Tehran lipid glucose study. Nutr Metab (Lond). 15(1):25.
  • Adachi H, Fujiwara Y, Ishii N. 1998. Effects of oxygen on protein carbonyl and aging in Caenorhabditis elegans mutants with long (age-1) and short (mev-1) life spans. J Gerontol A Biol Sci Med Sci. 53(4):B240–244.
  • Adam-Vizi V. 2005. Production of reactive oxygen species in brain mitochondria: contribution by electron transport chain and non-electron transport chain sources. Antioxid Redox Signal. 7(9–10):1140–1149.
  • Adesina SE, Kang BY, Bijli KM, Ma J, Cheng J, Murphy TC, Michael Hart C, Sutliff RL. 2015. Targeting mitochondrial reactive oxygen species to modulate hypoxia-induced pulmonary hypertension. Free Radic Biol Med. 87:36–47.
  • Adlam VJ, Harrison JC, Porteous CM, James AM, Smith RA, Murphy MP, Sammut IA. 2005. Targeting an antioxidant to mitochondria decreases cardiac ischemia-reperfusion injury. FASEB J. 19(9):1088–1095.
  • Agapova LS, Chernyak BV, Domnina LV, Dugina VB, Efimenko AY, Fetisova EK, Ivanova OY, Kalinina NI, Khromova NV, Kopnin BP, et al. 2008. Mitochondria-targeted plastoquinone derivatives as tools to interrupt execution of the aging program. 3. Inhibitory effect of SkQ1 on tumor development from p53-deficient cells. Biochemistry Mosc. 73(12):1300–1316.
  • Ahmed D, Roy D, Jaworski A, Edwards A, Abizaid A, Kumar A, Golshani A, Cassol E. 2019. Differential remodeling of the electron transport chain is required to support TLR3 and TLR4 signaling and cytokine production in macrophages. Sci Rep. 9(1):18801.
  • Ahmetov II, Naumov VA, Donnikov AE, Maciejewska-Karłowska A, Kostryukova ES, Larin AK, Maykova EV, Alexeev DG, Fedotovskaya ON, Generozov EV, et al. 2014. SOD2 gene polymorphism and muscle damage markers in elite athletes. Free Radic Res. 48(8):948–955.
  • Ahn J, Ambrosone CB, Kanetsky PA, Tian C, Lehman TA, Kropp S, Helmbold I, von Fournier D, Haase W, Sautter-Bihl ML, et al. 2006. Polymorphisms in genes related to oxidative stress (CAT, MnSOD, MPO, and eNOS) and acute toxicities from radiation therapy following lumpectomy for breast cancer. Clin Cancer Res. 12(23):7063–7070.
  • Akyol O, Yanik M, Elyas H, Namli M, Canatan H, Akin H, Yuce H, Yilmaz HR, Tutkun H, Sogut S, et al. 2005. Association between Ala-9Val polymorphism of Mn-SOD gene and schizophrenia. Prog Neuropsychopharmacol Biol Psychiatry. 29(1):123–131.
  • Al Hadithy AF, Ivanova SA, Pechlivanoglou P, Wilffert B, Semke A, Fedorenko O, Kornetova E, Ryadovaya L, Brouwers JR, Loonen AJ. 2010. Missense polymorphisms in three oxidative-stress enzymes (GSTP1, SOD2, and GPX1) and dyskinesias in Russian psychiatric inpatients from Siberia. Hum Psychopharmacol. 25(1):84–91.
  • Alam NM, Mills WC, Wong AA, Douglas RM, Szeto HH, Prusky GT. 2015. A mitochondrial therapeutic reverses visual decline in mouse models of diabetes. Dis Model Mech. 8(7):701–710.
  • Alassaf M, Daykin EC, Mathiaparanam J, Wolman MA. 2019. Pregnancy-associated plasma protein-aa supports hair cell survival by regulating mitochondrial function. Elife. 8:e47061.
  • Alfonso-Loeches S, Urena-Peralta JR, Morillo-Bargues MJ, Oliver-De La Cruz J, Guerri C. 2014. Role of mitochondria ROS generation in ethanol-induced NLRP3 inflammasome activation and cell death in astroglial cells. Front Cell Neurosci. 8:216.
  • Aljunaidy MM, Morton JS, Kirschenman R, Phillips T, Case CP, Cooke CM, Davidge ST. 2018. Maternal treatment with a placental-targeted antioxidant (MitoQ) impacts offspring cardiovascular function in a rat model of prenatal hypoxia. Pharmacol Res. 134:332–342.
  • Almeida M, Laurent MR, Dubois V, Claessens F, O’Brien CA, Bouillon R, Vanderschueren D, Manolagas SC. 2017. Estrogens and androgens in skeletal physiology and pathophysiology. Physiol Rev. 97(1):135–187.
  • Almomani R, Herkert JC, Posafalvi A, Post JG, Boven LG, van der Zwaag PA, Willems PHGM, van Veen-Hof IH, Verhagen JMA, Wessels MW, et al. 2020. Homozygous damaging SOD2 variant causes lethal neonatal dilated cardiomyopathy. J Med Genet. 57(1):23–30.
  • Al-Serri A, Anstee QM, Valenti L, Nobili V, Leathart JB, Dongiovanni P, Patch J, Fracanzani A, Fargion S, Day CP, et al. 2012. The SOD2 C47T polymorphism influences NAFLD fibrosis severity: evidence from case-control and intra-familial allele association studies. J Hepatol. 56(2):448–454.
  • Ambrosone CB, Ahn J, Singh KK, Rezaishiraz H, Furberg H, Sweeney C, Coles B, Trovato A. 2005. Polymorphisms in genes related to oxidative stress (MPO, MnSOD, CAT) and survival after treatment for breast cancer. Cancer Res. 65(3):1105–1111.
  • Ambrosone CB, Freudenheim JL, Thompson PA, Bowman E, Vena JE, Marshall JR, Graham S, Laughlin R, Nemoto T, Shields PG. 1999. Manganese superoxide dismutase (MnSOD) genetic polymorphisms, dietary antioxidants, and risk of breast cancer. Cancer Res. 59(3):602–606.
  • Amoedo ND, Dard L, Sarlak S, Mahfouf W, Blanchard W, Rousseau B, Izotte J, Claverol S, Lacombe D, Rezvani HR, et al. 2020. Targeting human lung adenocarcinoma with a supressor of mitochondrial superoxide production. Antioxid Redox Signal. 33(13):883–902.
  • Anderson EJ, Lustig ME, Boyle KE, Woodlief TL, Kane DA, Lin C-T, Price JW, Kang L, Rabinovitch PS, Szeto HH, et al. 2009. Mitochondrial H2O2 emission and cellular redox state link excess fat intake to insulin resistance in both rodents and humans. J Clin Invest. 119(3):573–581.
  • Andreassen OA, Ferrante RJ, Dedeoglu A, Albers DW, Klivenyi P, Carlson EJ, Epstein CJ, Beal MF. 2001. Mice with a partial deficiency of manganese superoxide dismutase show increased vulnerability to the mitochondrial toxins malonate, 3-nitropropionic acid, and MPTP. Exp Neurol. 167(1):189–195.
  • Andreassen OA, Ferrante RJ, Klivenyi P, Klein AM, Shinobu LA, Epstein CJ, Beal MF. 2000. Partial deficiency of manganese superoxide dismutase exacerbates a transgenic mouse model of amyotrophic lateral sclerosis. Ann Neurol. 47(4):447–455.
  • Andreev-Andrievskiy AA, Kolosova NG, Stefanova NA, Lovat MV, Egorov MV, Manskikh VN, Zinovkin RA, Galkin II, Prikhodko AS, Skulachev MV, et al. 2016. Efficacy of mitochondrial antioxidant plastoquinonyl-decyl-triphenylphosphonium bromide (SkQ1) in the rat model of autoimmune arthritis. Oxid Med Cell Longev. 2016:8703645.
  • Andreyev AY, Kushnareva YE, Murphy AN, Starkov AA. 2015. Mitochondrial ROS metabolism: 10 years later. Biochemistry Mosc. 80(5):517–531.
  • Andreyev AY, Kushnareva YE, Starkov AA. 2005. Mitochondrial metabolism of reactive oxygen species. Biochemistry Mosc. 70(2):200–214.
  • Anisimov VN, Bakeeva LE, Egormin PA, Filenko OF, Isakova EF, Manskikh VN, Mikhelson VM, Panteleeva AA, Pasyukova EG, Pilipenko DI, et al. 2008. Mitochondria-targeted plastoquinone derivatives as tools to interrupt execution of the aging program. 5. SkQ1 prolongs lifespan and prevents development of traits of senescence. Biochemistry Mosc. 73(12):1329–1342.
  • Anisimov VN, Egorov MV, Krasilshchikova MS, Lyamzaev KG, Manskikh VN, Moshkin MP, Novikov EA, Popovich IG, Rogovin KA, Shabalina IG, et al. 2011. Effects of the mitochondria-targeted antioxidant SkQ1 on lifespan of rodents. Aging (Albany NY). 3(11):1110–1119.
  • Ansher SS, Dolan P, Bueding E. 1983. Chemoprotective effects of two dithiolthiones and of butylhydroxyanisole against carbon tetrachloride and acetaminophen toxicity. Hepatology. 3(6):932–935.
  • Archer SL, Marsboom G, Kim GH, Zhang HJ, Toth PT, Svensson EC, Dyck JR, Gomberg-Maitland M, Thebaud B, Husain AN, et al. 2010. Epigenetic attenuation of mitochondrial superoxide dismutase 2 in pulmonary arterial hypertension: a basis for excessive cell proliferation and a new therapeutic target. Circulation. 121(24):2661–2671.
  • Arita Y, Harkness SH, Kazzaz JA, Koo HC, Joseph A, Melendez JA, Davis JM, Chander A, Li Y. 2006. Mitochondrial localization of catalase provides optimal protection from H2O2-induced cell death in lung epithelial cells. Am J Physiol Lung Cell Mol Physiol. 290(5):L978–986.
  • Arkat S, Umbarkar P, Singh S, Sitasawad SL. 2016. Mitochondrial peroxiredoxin-3 protects against hyperglycemia induced myocardial damage in diabetic cardiomyopathy. Free Radic Biol Med. 97:489–500.
  • Asayama K, Yokota S, Dobashi K, Hayashibe H, Kawaoi A, Nakazawa S. 1994. Purification and immunoelectron microscopic localization of cellular glutathione peroxidase in rat hepatocytes: quantitative analysis by postembedding method. Histochemistry. 102(3):213–219.
  • Ascencio-Montiel I. d J, Parra EJ, Valladares-Salgado A, Gómez-Zamudio JH, Kumate-Rodriguez J, Escobedo-de-la-Peña J, Cruz M. 2013. SOD2 gene Val16Ala polymorphism is associated with macroalbuminuria in Mexican type 2 diabetes patients: a comparative study and meta-analysis. BMC Med Genet. 14(1):110.
  • Asimakis GK, Lick S, Patterson C. 2002. Postischemic recovery of contractile function is impaired in SOD2(+/−) but not SOD1(+/−) mouse hearts. Circulation. 105(8):981–986.
  • Asin-Cayuela J, Manas AR, James AM, Smith RA, Murphy MP. 2004. Fine-tuning the hydrophobicity of a mitochondria-targeted antioxidant. FEBS Lett. 571(1–3):9–16.
  • Atilgan D, Parlaktas BS, Uluocak N, Kolukcu E, Erdemir F, Ozyurt H, Erkorkmaz U. 2014. The relationship between ALA16VAL single gene polymorphism and renal cell carcinoma. Adv Urol. 2014:1–5.
  • Aynali G, Dogan M, Sutcu R, Yuksel O, Yariktas M, Unal F, Yasan H, Ceyhan B, Tuz M. 2013. Polymorphic variants of MnSOD Val16Ala, CAT-262 C<T and GPx1 Pro198Leu genotypes and the risk of laryngeal cancer in a smoking population. J Laryngol Otol. 127(10):997–1000.
  • Babizhayev MA. 2016. Generation of reactive oxygen species in the anterior eye segment. Synergistic codrugs of N-acetylcarnosine lubricant eye drops and mitochondria-targeted antioxidant act as a powerful therapeutic platform for the treatment of cataracts and primary open-angle glaucoma. Biochim Biophys Acta Clin. 6:49–68.
  • Bae SH, Sung SH, Lee HE, Kang HT, Lee SK, Oh SY, Woo HA, Kil IS, Rhee SG. 2012. Peroxiredoxin III and sulfiredoxin together protect mice from pyrazole-induced oxidative liver injury. Antioxid Redox Signal. 17(10):1351–1361.
  • Baek JY, Park S, Park J, Jang JY, Wang SB, Kim SR, Woo HA, Lim KM, Chang TS. 2017. Protective role of mitochondrial peroxiredoxin III against UVB-induced apoptosis of epidermal keratinocytes. J Invest Dermatol. 137(6):1333–1342.
  • Bag A, Bag N. 2008. Target sequence polymorphism of human manganese superoxide dismutase gene and its association with cancer risk: a review. Cancer Epidemiol Biomarkers Prev. 17(12):3298–3305.
  • Bai J, Cederbaum AI. 2001. Adenovirus-mediated overexpression of catalase in the cytosolic or mitochondrial compartment protects against cytochrome P450 2E1-dependent toxicity in HepG2 cells. J Biol Chem. 276(6):4315–4321.
  • Bakeeva LE, Barskov IV, Egorov MV, Isaev NK, Kapelko VI, Kazachenko AV, Kirpatovsky VI, Kozlovsky SV, Lakomkin VL, Levina SB, et al. 2008. Mitochondria-targeted plastoquinone derivatives as tools to interrupt execution of the aging program. 2. Treatment of some ROS- and age-related diseases (heart arrhythmia, heart infarctions, kidney ischemia, and stroke). Biochemistry Mosc. 73(12):1288–1299.
  • Bakker PR, van Harten PN, van Os J. 2008. Antipsychotic-induced tardive dyskinesia and polymorphic variations in COMT, DRD2, CYP1A2 and MnSOD genes: a meta-analysis of pharmacogenetic interactions. Mol Psychiatry. 13(5):544–556.
  • Balaban RS, Nemoto S, Finkel T. 2005. Mitochondria, oxidants, and aging. Cell. 120(4):483–495.
  • Baldari S, Di Rocco G, Trivisonno A, Samengo D, Pani G, Toietta G. 2016. Promotion of survival and engraftment of transplanted adipose tissue-derived stromal and vascular cells by overexpression of manganese superoxide dismutase. Int J Mol Sci. 17(7):1082.
  • Ballinger SW, Patterson C, Knight-Lozano CA, Burow DL, Conklin CA, Hu Z, Reuf J, Horaist C, Lebovitz R, Hunter GC, et al. 2002. Mitochondrial integrity and function in atherogenesis. Circulation. 106(5):544–549.
  • Banescu C, Trifa AP, Voidazan S, Moldovan VG, Macarie I, Benedek Lazar E, Dima D, Duicu C, Dobreanu M. 2014. CAT, GPX1, MnSOD, GSTM1, GSTT1, and GSTP1 genetic polymorphisms in chronic myeloid leukemia: a case-control study. Oxid Med Cell Longev. 2014:875861.
  • Bartell LR, Fortier LA, Bonassar LJ, Szeto HH, Cohen I, Delco ML. 2020. Mitoprotective therapy prevents rapid, strain-dependent mitochondrial dysfunction after articular cartilage injury. J Orthop Res. 38(6):1257–1267.
  • Bartell SM, Kim HN, Ambrogini E, Han L, Iyer SS, Ucer S, Rabinovitch P, Jilka RL, Weinstein RS, Zhao H, et al. 2014. FoxO proteins restrain osteoclastogenesis and bone resorption by attenuating H2O2 accumulation. Nat Commun. 5:3773.
  • Batinic-Haberle I, Tovmasyan A, Spasojevic I. 2018. Mn porphyrin-based redox-active drugs: differential effects as cancer therapeutics and protectors of normal tissue against oxidative injury. Antioxid Redox Signal. 29(16):1691–1724.
  • Bayeva M, Gheorghiade M, Ardehali H. 2013. Mitochondria as a therapeutic target in heart failure. J Am Coll Cardiol. 61(6):599–610.
  • Becer E, Çırakoğlu A. 2015. Association of the Ala16Val MnSOD gene polymorphism with plasma leptin levels and oxidative stress biomarkers in obese patients. Gene. 568(1):35–39.
  • Ben-Zaken S, Eliakim A, Nemet D, Kassem E, Meckel Y. 2013. Increased prevalence of MnSOD genetic polymorphism in endurance and power athletes. Free Radic Res. 47(12):1002–1008.
  • Bergman M, Ahnstrom M, Palmeback Wegman P, Wingren S. 2005. Polymorphism in the manganese superoxide dismutase (MnSOD) gene and risk of breast cancer in young women. J Cancer Res Clin Oncol. 131(7):439–444.
  • Bewick MA, Conlon MS, Lafrenie RM. 2008. Polymorphisms in manganese superoxide dismutase, myeloperoxidase and glutathione-S-transferase and survival after treatment for metastatic breast cancer. Breast Cancer Res Treat. 111(1):93–101.
  • Bica CG, de Moura da Silva LL, Toscani NV, da Cruz IB, Sa G, Graudenz MS, Zettler CG. 2009. MnSOD gene polymorphism association with steroid-dependent cancer. Pathol Oncol Res. 15(1):19–24.
  • Biosa A, Sanchez-Martinez A, Filograna R, Terriente-Felix A, Alam SM, Beltramini M, Bubacco L, Bisaglia M, Whitworth AJ. 2018. Superoxide dismutating molecules rescue the toxic effects of PINK1 and parkin loss. Hum Mol Genet. 27(9):1618–1629.
  • Birk AV, Liu S, Soong Y, Mills W, Singh P, Warren JD, Seshan SV, Pardee JD, Szeto HH. 2013. The mitochondrial-targeted compound SS-31 re-energizes ischemic mitochondria by interacting with cardiolipin. J Am Soc Nephrol. 24(8):1250–1261.
  • Boden MJ, Brandon AE, Tid-Ang JD, Preston E, Wilks D, Stuart E, Cleasby ME, Turner N, Cooney GJ, Kraegen EW. 2012. Overexpression of manganese superoxide dismutase ameliorates high-fat diet-induced insulin resistance in rat skeletal muscle. Am J Physiol Endocrinol Metab. 303(6):E798–805.
  • Bond ST, Kim J, Calkin AC, Drew BG. 2019. The antioxidant moiety of MitoQ imparts minimal metabolic effects in adipose tissue of high fat fed mice. Front Physiol. 10:543.
  • Bonetta R. 2018. Potential therapeutic applications of MnSODs and SOD-mimetics. Chemistry. 24(20):5032–5041.
  • Bordt EA, Polster BM. 2014. NADPH oxidase- and mitochondria-derived reactive oxygen species in proinflammatory microglial activation: a bipartisan affair? Free Radic Biol Med. 76:34–46.
  • Boroumand F, Mahmoudinasab H, Saadat M. 2018. Association of the SOD2 (rs2758339 and rs5746136) polymorphisms with the risk of heroin dependency and the SOD2 expression levels. Gene. 649:27–31.
  • Boskovic M, Vovk T, Saje M, Goricar K, Dolzan V, Kores Plesnicar B, Grabnar I. 2013. Association of SOD2, GPX1, CAT, and TNF genetic polymorphisms with oxidative stress, neurochemistry, psychopathology, and extrapyramidal symptoms in schizophrenia. Neurochem Res. 38(2):433–442.
  • Botre C, Shahu A, Adkar N, Shouche Y, Ghaskadbi S, Ashma R. 2015. Superoxide dismutase 2 polymorphisms and osteoporosis in Asian Indians: a genetic association analysis. Cell Mol Biol Lett. 20(4):685–697.
  • Boucard A, Marin F, Monternier PA, Brand M, Le Grand B. 2019. A specific complex I-induced ROS modulator, OP2113, is a new cardioproctective agent against acute myocardial infarction injuries during reperfusion. Eur Heart J. 40:2818.
  • Bouitbir J, Singh F, Charles AL, Schlagowski AI, Bonifacio A, Echaniz-Laguna A, Geny B, Krahenbuhl S, Zoll J. 2016. Statins trigger mitochondrial reactive oxygen species-induced apoptosis in glycolytic skeletal muscle. Antioxid Redox Signal. 24(2):84–98.
  • Boveris A, Chance B. 1973. The mitochondrial generation of hydrogen peroxide. General properties and effect of hyperbaric oxygen. Biochem J. 134(3):707–716.
  • Braakhuis AJ, Nagulan R, Somerville V. 2018. The effect of MitoQ on aging-related biomarkers: a systematic review and meta-analysis. Oxid Med Cell Longev. 2018:8575263
  • Brand MD. 2010. The sites and topology of mitochondrial superoxide production. Exp Gerontol. 45(7–8):466–472.
  • Brand MD. 2016. Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radic Biol Med. 100:14–31.
  • Brand MD, Affourtit C, Esteves TC, Green K, Lambert AJ, Miwa S, Pakay JL, Parker N. 2004. Mitochondrial superoxide: production, biological effects, and activation of uncoupling proteins. Free Radic Biol Med. 37(6):755–767.
  • Brand MD, Goncalves RLS, Orr AL, Vargas L, Gerencser AA, Borch Jensen M, Wang YT, Melov S, Turk CN, Matzen JT, et al. 2016. Suppressors of superoxide-H2O2 production at site IQ of mitochondrial Complex I protect against stem cell hyperplasia and ischemia-reperfusion injury. Cell Metab. 24(4):582–592.
  • Bronner DN, Abuaita BH, Chen X, Fitzgerald KA, Nunez G, He Y, Yin XM, O’Riordan MX. 2015. Endoplasmic reticulum stress activates the inflammasome via NLRP3- and caspase-2-driven mitochondrial damage. Immunity. 43(3):451–462.
  • Brookes PS. 2005. Mitochondrial H + leak and ROS generation: an odd couple. Free Radic Biol Med. 38(1):12–23.
  • Brookes PS, Yoon Y, Robotham JL, Anders MW, Sheu SS. 2004. Calcium, ATP, and ROS: a mitochondrial love-hate triangle. Am J Physiol Cell Physiol. 287(4):C817–833.
  • Brose RD, Avramopoulos D, Smith KD. 2012. SOD2 as a potential modifier of X-linked adrenoleukodystrophy clinical phenotypes. J Neurol. 259(7):1440–1447.
  • Brown GC, Borutaite V. 2012. There is no evidence that mitochondria are the main source of reactive oxygen species in mammalian cells. Mitochondrion. 12(1):1–4.
  • Brown KA, Didion SP, Andresen JJ, Faraci FM. 2007. Effect of aging, MnSOD deficiency, and genetic background on endothelial function: evidence for MnSOD haploinsufficiency. Arterioscler Thromb Vasc Biol. 27(9):1941–1946.
  • Brown DA, Hale SL, Baines CP, del Rio CL, Hamlin RL, Yueyama Y, Kijtawornrat A, Yeh ST, Frasier CR, Stewart LM, et al. 2014. Reduction of early reperfusion injury with the mitochondria-targeting peptide bendavia. J Cardiovasc Pharmacol Ther. 19(1):121–132.
  • Brown AL, Lupo PJ, Okcu MF, Lau CC, Rednam S, Scheurer ME. 2015. SOD2 genetic variant associated with treatment-related ototoxicity in cisplatin-treated pediatric medulloblastoma. Cancer Med. 4(11):1679–1686.
  • Brown E, Ash JD, Lewin AS. 2017. Deletion of mitochondrial antioxidant enzyme Sod2 induces light-dependent retinal degeneration with aging. ARVO (the Association for Research in Vision and Ophthalmology). Invest Ophthalmol Vis Sci. 58(8):2294.
  • Brunelle JK, Bell EL, Quesada NM, Vercauteren K, Tiranti V, Zeviani M, Scarpulla RC, Chandel NS. 2005. Oxygen sensing requires mitochondrial ROS but not oxidative phosphorylation. Cell Metab. 1(6):409–414.
  • Bruton JD, Place N, Yamada T, Silva JP, Andrade FH, Dahlstedt AJ, Zhang SJ, Katz A, Larsson NG, Westerblad H. 2008. Reactive oxygen species and fatigue-induced prolonged low-frequency force depression in skeletal muscle fibres of rats, mice and SOD2 overexpressing mice. J Physiol (Lond). 586(1):175–184.
  • Brzheskiy VV, Efimova EL, Vorontsova TN, Alekseev VN, Gusarevich OG, Shaidurova KN, Ryabtseva AA, Andryukhina OM, Kamenskikh TG, Sumarokova ES, et al. 2015. Results of a multicenter, randomized, double-masked, placebo-controlled clinical study of the efficacy and safety of visomitin eye drops in patients with dry eye syndrome. Adv Ther. 32(12):1263–1279.
  • Bullock JJ, Mehta SL, Lin Y, Lolla P, Li PA. 2009. Hyperglycemia-enhanced ischemic brain damage in mutant manganese SOD mice is associated with suppression of HIF-1alpha. Neurosci Lett. 456(2):89–92.
  • Bulua AC, Simon A, Maddipati R, Pelletier M, Park H, Kim KY, Sack MN, Kastner DL, Siegel RM. 2011. Mitochondrial reactive oxygen species promote production of proinflammatory cytokines and are elevated in TNFR1-associated periodic syndrome (TRAPS). J Exp Med. 208(3):519–533.
  • Burbulla LF, Song P, Mazzulli JR, Zampese E, Wong YC, Jeon S, Santos DP, Blanz J, Obermaier CD, Strojny C, et al. 2017. Dopamine oxidation mediates mitochondrial and lysosomal dysfunction in Parkinson’s disease. Science. 357(6357):1255–1261.
  • Burri RJ, Stock RG, Cesaretti JA, Atencio DP, Peters S, Peters CA, Fan G, Stone NN, Ostrer H, Rosenstein BS. 2008. Association of single nucleotide polymorphisms in SOD2, XRCC1 and XRCC3 with susceptibility for the development of adverse effects resulting from radiotherapy for prostate cancer. Radiat Res. 170(1):49–59.
  • Buskiewicz IA, Montgomery T, Yasewicz EC, Huber SA, Murphy MP, Hartley RC, Kelly R, Crow MK, Perl A, Budd RC, et al. 2016. Reactive oxygen species induce virus-independent MAVS oligomerization in systemic lupus erythematosus. Sci Signal. 9(456):ra115.
  • Cabreiro F, Ackerman D, Doonan R, Araiz C, Back P, Papp D, Braeckman BP, Gems D. 2011. Increased life span from overexpression of superoxide dismutase in Caenorhabditis elegans is not caused by decreased oxidative damage. Free Rad Biol Med. 51(8):1575–1582.
  • Cadenas E. 2004. Mitochondrial free radical production and cell signaling. Mol Aspects Med. 25(1–2):17–26.
  • Cadenas E, Davies KJ. 2000. Mitochondrial free radical generation, oxidative stress, and aging. Free Radic Biol Med. 29(3–4):222–230.
  • Cai J, Jiang Y, Zhang M, Zhao H, Li H, Li K, Zhang X, Qiao T. 2018. Protective effects of mitochondrion-targeted peptide SS-31 against hind limb ischemia-reperfusion injury. J Physiol Biochem. 74(2):335–343.
  • Cai M, Li J, Lin S, Chen X, Huang J, Jiang X, Yang L, Luo Y. 2015. Mitochondria-targeted antioxidant peptide SS31 protects cultured human lens epithelial cells against oxidative stress. Curr Eye Res. 40(8):822–829.
  • Cai Q, Shu XO, Wen W, Cheng JR, Dai Q, Gao YT, Zheng W. 2004. Genetic polymorphism in the manganese superoxide dismutase gene, antioxidant intake, and breast cancer risk: results from the Shanghai Breast Cancer Study. Breast Cancer Res. 6(6):R647–655.
  • Calkins MJ, Manczak M, Mao P, Shirendeb U, Reddy PH. 2011. Impaired mitochondrial biogenesis, defective axonal transport of mitochondria, abnormal mitochondrial dynamics and synaptic degeneration in a mouse model of Alzheimer’s disease. Hum Mol Genet. 20(23):4515–4529.
  • Callio J, Oury TD, Chu CT. 2005. Manganese superoxide dismutase protects against 6-hydroxydopamine injury in mouse brains. J Biol Chem. 280(18):18536–18542.
  • Campbell MD, Duan J, Samuelson AT, Gaffrey MJ, Merrihew GE, Egertson JD, Wang L, Bammler TK, Moore RJ, White CC, et al. 2019. Improving mitochondrial function with SS-31 reverses age-related redox stress and improves exercise tolerance in aged mice. Free Radic Biol Med. 134:268–281.
  • Cantarero A, Alonso-Alvarez C. 2017. Mitochondria-targeted molecules determine the redness of the zebra finch bill. Biol Lett. 13(10):20170455.
  • Canugovi C, Stevenson MD, Vendrov AE, Hayami T, Robidoux J, Xiao H, Zhang YY, Eitzman DT, Runge MS, Madamanchi NR. 2019. Increased mitochondrial NADPH oxidase 4 (NOX4) expression in aging is a causative factor in aortic stiffening. Redox Biol. 26:101288.
  • Carvalho C, Katz PS, Dutta S, Katakam PV, Moreira PI, Busija DW. 2014. Increased susceptibility to amyloid-beta toxicity in rat brain microvascular endothelial cells under hyperglycemic conditions. J Alzheimers Dis. 38(1):75–83.
  • Case AJ, McGill JL, Tygrett LT, Shirasawa T, Spitz DR, Waldschmidt TJ, Legge KL, Domann FE. 2011. Elevated mitochondrial superoxide disrupts normal T cell development, impairing adaptive immune responses to an influenza challenge. Free Rad Biol Med. 50(3):448–458.
  • Castro L, Tortora V, Mansilla S, Radi R. 2019. Aconitases: non-redox iron-sulfur proteins sensitive to reactive species. Acc Chem Res. 52(9):2609–2619.
  • Celojevic D, Nilsson S, Behndig A, Tasa G, Juronen E, Karlsson JO, Zetterberg H, Petersen A, Zetterberg M. 2013. Superoxide dismutase gene polymorphisms in patients with age-related cataract. Ophthalmic Genet. 34(3):140–145.
  • Celojevic D, Nilsson S, Kalaboukhova L, Tasa G, Juronen E, Sjolander A, Zetterberg H, Zetterberg M. 2014. Genetic variation of superoxide dismutases in patients with primary open-angle glaucoma. Ophthalmic Genet. 35(2):79–84.
  • Chacko BK, Reily C, Srivastava A, Johnson MS, Ye Y, Ulasova E, Agarwal A, Zinn KR, Murphy MP, Kalyanaraman B, et al. 2010. Prevention of diabetic nephropathy in Ins2(+/)(-)(AkitaJ) mice by the mitochondria-targeted therapy MitoQ. Biochem J. 432(1):9–19.
  • Chacko BK, Srivastava A, Johnson MS, Benavides GA, Chang MJ, Ye Y, Jhala N, Murphy MP, Kalyanaraman B, Darley-Usmar VM. 2011. Mitochondria-targeted ubiquinone (MitoQ) decreases ethanol-dependent micro and macro hepatosteatosis. Hepatology. 54(1):153–163.
  • Chae HZ, Kim HJ, Kang SW, Rhee SG. 1999. Characterization of three isoforms of mammalian peroxiredoxin that reduce peroxides in the presence of thioredoxin. Diabetes Res Clin Pract. 45(2–3):101–112.
  • Chaiswing L, Cole MP, Ittarat W, Szweda LI, St Clair DK, Oberley TD. 2005. Manganese superoxide dismutase and inducible nitric oxide synthase modify early oxidative events in acute adriamycin-induced mitochondrial toxicity. Mol Cancer Ther. 4(7):1056–1064.
  • Chan JM, Oh WK, Xie W, Regan MM, Stampfer MJ, King IB, Abe M, Kantoff PW. 2009. Plasma selenium, manganese superoxide dismutase, and intermediate- or high-risk prostate cancer. J Clin Oncol. 27(22):3577–3583.
  • Chance B, Sies H, Boveris A. 1979. Hydroperoxide metabolism in mammalian organs. Physiol Rev. 59(3):527–605.
  • Chandran K, Aggarwal D, Migrino RQ, Joseph J, McAllister D, Konorev EA, Antholine WE, Zielonka J, Srinivasan S, Avadhani NG, et al. 2009. Doxorubicin inactivates myocardial cytochrome c oxidase in rats: cardioprotection by Mito-Q. Biophys J. 96(4):1388–1398.
  • Chang TS, Cho CS, Park S, Yu S, Kang SW, Rhee SG. 2004. Peroxiredoxin III, a mitochondrion-specific peroxidase, regulates apoptotic signaling by mitochondria. J Biol Chem. 279(40):41975–41984.
  • Chao CT, Chen YC, Chiang CK, Huang JW, Fang CC, Chang CC, Yen CJ. 2016. Interplay between superoxide dismutase, glutathione peroxidase, and peroxisome proliferator activated receptor gamma polymorphisms on the risk of end-stage renal disease among Han Chinese patients. Oxid Med Cell Longev. 2016:8516748.
  • Chelombitko MA, Averina OA, Vasil’eva TV, Dvorianinova EE, Egorov MV, Pletjushkina OY, Popova EN, Fedorov AV, Romashchenko VP, Ilyinskaya OP. 2017. Comparison of the effects of mitochondria-targeted antioxidant 10-(6’-plastoquinonyl)decyltriphenylphosphonium bromide (SkQ1) and a fragment of its molecule dodecyltriphenylphosphonium on carrageenan-induced acute inflammation in mouse model of subcuteneous air pouch. Bull Exp Biol Med. 162(6):730–733.
  • Chen PM, Cheng YW, Wu TC, Chen CY, Lee H. 2015. MnSOD overexpression confers cisplatin resistance in lung adenocarcinoma via the NF-kappaB/Snail/Bcl-2 pathway. Free Radic Biol Med. 79:127–137.
  • Chen K, Dai H, Yuan J, Chen J, Lin L, Zhang W, Wang L, Zhang J, Li K, He Y. 2018. Optineurin-mediated mitophagy protects renal tubular epithelial cells against accelerated senescence in diabetic nephropathy. Cell Death Dis. 9(2):105.
  • Chen L, Na R, Gu M, Salmon AB, Liu Y, Liang H, Qi W, Van Remmen H, Richardson A, Ran Q. 2008. Reduction of mitochondrial H2O2 by overexpressing peroxiredoxin 3 improves glucose tolerance in mice. Aging Cell. 7(6):866–878.
  • Chen L, Na R, Ran Q. 2014. Enhanced defense against mitochondrial hydrogen peroxide attenuates age-associated cognition decline. Neurobiol Aging. 35(11):2552–2561.
  • Chen Z, Siu B, Ho YS, Vincent R, Chua CC, Hamdy RC, Chua BH. 1998. Overexpression of MnSOD protects against myocardial ischemia/reperfusion injury in transgenic mice. J Mol Cell Cardiol. 30(11):2281–2289.
  • Chen S, Wang Y, Zhang H, Chen R, Lv F, Li Z, Jiang T, Lin D, Zhang H, Yang L, et al. 2019. The antioxidant MitoQ protects against CSE-induced endothelial barrier injury and inflammation by inhibiting ROS and autophagy in human umbilical vein endothelial cells. Int J Biol Sci. 15(7):1440–1451.
  • Chen L, Yoo SE, Na R, Liu Y, Ran Q. 2012. Cognitive impairment and increased Abeta levels induced by paraquat exposure are attenuated by enhanced removal of mitochondrial H(2)O(2). Neurobiol Aging. 33(2):432 e415.
  • Chen FQ, Zheng HW, Schacht J, Sha SH. 2013. Mitochondrial peroxiredoxin 3 regulates sensory cell survival in the cochlea. PloS One. 8(4):e61999.
  • Chen YR, Zweier JL. 2014. Cardiac mitochondria and reactive oxygen species generation. Circ Res. 114(3):524–537.
  • Cheng YJ, Wang YD, Liu Q, Dong ZM, Wu FP, Yang XR, Qiao XY, Zhang J. 2009. Association of single nucleotide polymorphism of MnSOD gene with carcinogenesis and development of esophageal squamous cell carcinoma. Zhonghua Zhong Liu Za Zhi. 31(11):831–835.
  • Chen W, Guo C, Jia Z, Wang J, Xia M, Li C, Li M, Yin Y, Tang X, Chen T, et al. 2020. Inhibition of mitochondrial ROS by mitoQ alleviates white matter injury and improves outcomes after intracerebral haemorrhage in mice. Oxid Med Cell Longev. 2020:8285065.
  • Chen H, Yu M, Li M, Zhao R, Zhu Q, Zhou W, Lu M, Lu Y, Zheng T, Jiang J, et al. 2012. Polymorphic variations in manganese superoxide dismutase (MnSOD), glutathione peroxidase-1 (GPX1), and catalase (CAT) contribute to elevated plasma triglyceride levels in Chinese patients with type 2 diabetes or diabetic cardiovascular disease. Mol Cell Biochem. 363(1–2):85–91.
  • Chistyakov DA, Savost’anov KV, Zotova EV, Nosikov VV. 2001. Polymorphisms in the Mn-SOD and EC-SOD genes and their relationship to diabetic neuropathy in type 1 diabetes mellitus. BMC Med Genet. 2(1):4.
  • Chmielewski NN, Caressi C, Giedzinski E, Parihar VK, Limoli CL. 2016. Contrasting the effects of proton irradiation on dendritic complexity of subiculum neurons in wild type and MCAT mice. Environ Mol Mutagen. 57(5):364–371.
  • Cho SJ, Park JW, Kang JS, Kim WH, Juhnn YS, Lee JS, Kim YH, Ko YS, Nam SY, Lee BL. 2008. Nuclear factor-kappaB dependency of doxorubicin sensitivity in gastric cancer cells is determined by manganese superoxide dismutase expression. Cancer Sci. 99(6):1117–1124.
  • Cho S, Szeto HH, Kim E, Kim H, Tolhurst AT, Pinto JT. 2007. A novel cell-permeable antioxidant peptide, SS31, attenuates ischemic brain injury by down-regulating CD36. J Biol Chem. 282(7):4634–4642.
  • Cho J, Won K, Wu D, Soong Y, Liu S, Szeto HH, Hong MK. 2007. Potent mitochondria-targeted peptides reduce myocardial infarction in rats. Coron Artery Dis. 18(3):215–220.
  • Choi JY, Neuhouser ML, Barnett MJ, Hong CC, Kristal AR, Thornquist MD, King IB, Goodman GE, Ambrosone CB. 2008. Iron intake, oxidative stress-related genes (MnSOD and MPO) and prostate cancer risk in CARET cohort. Carcinogenesis. 29(5):964–970.
  • Chouchani ET, Kazak L, Jedrychowski MP, Lu GZ, Erickson BK, Szpyt J, Pierce KA, Laznik-Bogoslavski D, Vetrivelan R, Clish CB, et al. 2016. Mitochondrial ROS regulate thermogenic energy expenditure and sulfenylation of UCP1. Nature. 532(7597):112–116.
  • Chouchani ET, Pell VR, James AM, Work LM, Saeb-Parsy K, Frezza C, Krieg T, Murphy MP. 2016. A unifying mechanism for mitochondrial superoxide production during ischemia-reperfusion injury. Cell Metab. 23(2):254–263.
  • Chowdhury A, Aich A, Jain G, Wozny K, Luchtenborg C, Hartmann M, Bernhard O, Balleiniger M, Alfar EA, Zieseniss A, et al. 2018. Defective mitochondrial cardiolipin remodeling dampens HIF-1alpha expression in hypoxia. Cell Rep. 25(3):561–570 e566.
  • Chu FF, Esworthy RS, Doroshow JH, Grasberger H, Donko A, Leto TL, Gao Q, Shen B. 2017. Deficiency in Duox2 activity alleviates ileitis in GPx1- and GPx2-knockout mice without affecting apoptosis incidence in the crypt epithelium. Redox Biol. 11:144–156.
  • Chua PJ, Lee EH, Yu Y, Yip GW, Tan PH, Bay BH. 2010. Silencing the Peroxiredoxin III gene inhibits cell proliferation in breast cancer. Int J Oncol. 36(2):359–364.
  • Church SL, Grant JW, Ridnour LA, Oberley LW, Swanson PE, Meltzer PS, Trent JM. 1993. Increased manganese superoxide dismutase expression suppresses the malignant phenotype of human melanoma cells. Proc Natl Acad Sci USA. 90(7):3113–3117.
  • Cline JM, Dugan G, Bourland JD, Perry DL, Stitzel JD, Weaver AA, Jiang C, Tovmasyan A, Owzar K, Spasojevic I, et al. 2018. Post-irradiation treatment with a superoxide dismutase mimic, MnTnHex-2-PyP(5+), mitigates radiation injury in the lungs of non-human primates after whole-thorax exposure to ionizing radiation. Antioxidants (Basel). 7(3):40.
  • Collin F. 2019. Chemical basis of reactive oxygen species reactivity and involvement in neurodegenerative diseases. Int J Mol Sci. 20(10):2407.
  • Collins JA, Wood ST, Nelson KJ, Rowe MA, Carlson CS, Chubinskaya S, Poole LB, Furdui CM, Loeser RF. 2016. Oxidative stress promotes peroxiredoxin hyperoxidation and attenuates pro-survival signaling in aging chondrocytes. J Biol Chem. 291(13):6641–6654.
  • Connor KM, Hempel N, Nelson KK, Dabiri G, Gamarra A, Belarmino J, Van De Water L, Mian BM, Melendez JA. 2007. Manganese superoxide dismutase enhances the invasive and migratory activity of tumor cells. Cancer Res. 67(21):10260–10267.
  • Constantinou C, Apidianakis Y, Psychogios N, Righi V, Mindrinos MN, Khan N, Swartz HM, Szeto HH, Tompkins RG, Rahme LG, et al. 2016. In vivo high-resolution magic angle spinning magnetic and electron paramagnetic resonance spectroscopic analysis of mitochondria-targeted peptide in Drosophila melanogaster with trauma-induced thoracic injury. Int J Mol Med. 37(2):299–308.
  • Coskun C, Verim A, Farooqi AA, Turan S, Mezani B, Kucukhuseyin O, Tolgahan Hakan M, Ergen A, Yaylim I. 2016. Are there possible associations between MnSOD and GPx1 gene variants for laryngeal cancer risk or disease progression? Cell Mol Biol (Noisy-le-Grand). 62(5):25–30.
  • Costa Pereira C, Duraes C, Coelho R, Gracio D, Silva M, Peixoto A, Lago P, Pereira M, Catarino T, Pinho S, et al. 2017. Association between polymorphisms in antioxidant genes and inflammatory bowel disease. PLoS One. 12(1):e0169102.
  • Coudray C, Fouret G, Lambert K, Ferreri C, Rieusset J, Blachnio-Zabielska A, Lecomte J, Ebabe Elle R, Badia E, Murphy MP, et al. 2016. A mitochondrial-targeted ubiquinone modulates muscle lipid profile and improves mitochondrial respiration in obesogenic diet-fed rats. Br J Nutr. 115(7):1155–1166.
  • Cox CS, McKay SE, Holmbeck MA, Christian BE, Scortea AC, Tsay AJ, Newman LE, Shadel GS. 2018. Mitohormesis in mice via sustained basal activation of mtochondrial and antioxidant signaling. Cell Metab. 28(5):776–786 e775.
  • Cox DG, Tamimi RM, Hunter DJ. 2006. Gene x Gene interaction between MnSOD and GPX-1 and breast cancer risk: a nested case-control study. BMC Cancer. 6(1):217.
  • Cox AG, Winterbourn CC, Hampton MB. 2009. Mitochondrial peroxiredoxin involvement in antioxidant defence and redox signalling. Biochem J. 425(2):313–325.
  • Craven PA, Phillips SL, Melhem MF, Liachenko J, DeRubertis FR. 2001. Overexpression of manganese superoxide dismutase suppresses increases in collagen accumulation induced by culture of mesangial cells in high-media glucose. Metab Clin Exp. 50(9):1043–1048.
  • Crawford A, Fassett RG, Coombes JS, Kunde DA, Ahuja KD, Robertson IK, Ball MJ, Geraghty DP. 2011. Glutathione peroxidase, superoxide dismutase and catalase genotypes and activities and the progression of chronic kidney disease. Nephrol Dial Transplant. 26(9):2806–2813.
  • Crooks DR, Maio N, Lane AN, Jarnik M, Higashi RM, Haller RG, Yang Y, Fan TW, Linehan WM, Rouault TA. 2018. Acute loss of iron-sulfur clusters results in metabolic reprogramming and generation of lipid droplets in mammalian cells. J Biol Chem. 293(21):8297–8311.
  • Csiszar A, Yabluchanskiy A, Ungvari A, Ungvari Z, Tarantini S. 2019. Overexpression of catalase targeted to mitochondria improves neurovascular coupling responses in aged mice. Geroscience. 41(5):609–617.
  • Cui X, Lu X, Hiura M, Oda M, Miyazaki W, Katoh T. 2013. Evaluation of genetic polymorphisms in patients with multiple chemical sensitivity. PLoS One. 8(8):e73708.
  • Cumurcu BE, Ozyurt H, Ates O, Gul IG, Demir S, Karlidag R. 2013. Analysis of manganese superoxide dismutase (MnSOD: Ala-9Val) and glutathione peroxidase (GSH-Px: Pro 197 Leu) gene polymorphisms in mood disorders. Bosn J Basic Med Sci. 13(2):109–113.
  • Curtis C, Landis GN, Folk D, Wehr NB, Hoe N, Waskar M, Abdueva D, Skvortsov D, Ford D, Luu A, et al. 2007. Transcriptional profiling of MnSOD-mediated lifespan extension in Drosophila reveals a species-general network of aging and metabolic genes. Genome Biol. 8(12):R262.
  • Czigler A, Toth L, Szarka N, Berta G, Amrein K, Czeiter E, Lendvai-Emmert D, Bodo K, Tarantini S, Koller A, et al. 2019. Hypertension exacerbates cerebrovascular oxidative stress induced by mild traumatic rrain injury: protective effects of the mitochondria-targeted antioxidative peptide SS-31. J Neurotrauma. 36(23):3309–3315.
  • D’Souza AD, Parish IA, Krause DS, Kaech SM, Shadel GS. 2013. Reducing mitochondrial ROS improves disease-related pathology in a mouse model of ataxia-telangiectasia. Mol Ther. 21(1):42–48.
  • da Cruz Jung IE, da Cruz IBM, Barbisan F, Trott A, Houenou LJ, Osmarin Turra B, Duarte T, de Souza Praia R, Maia-Ribeiro EA, da Costa Escobar Piccoli J, et al. 2020. Superoxide imbalance triggered by Val16Ala-SOD2 polymorphism increases the risk of depression and self-reported psychological stress in free-living elderly people. Mol Genet Genomic Med. 8(2):e1080.
  • Dai DF, Chen T, Szeto H, Nieves-Cintron M, Kutyavin V, Santana LF, Rabinovitch PS. 2011. Mitochondrial targeted antioxidant peptide ameliorates hypertensive cardiomyopathy. J Am Coll Cardiol. 58(1):73–82.
  • Dai DF, Chen T, Wanagat J, Laflamme M, Marcinek DJ, Emond MJ, Ngo CP, Prolla TA, Rabinovitch PS. 2010. Age-dependent cardiomyopathy in mitochondrial mutator mice is attenuated by overexpression of catalase targeted to mitochondria. Aging Cell. 9(4):536–544.
  • Dai DF, Chiao YA, Marcinek DJ, Szeto HH, Rabinovitch PS. 2014. Mitochondrial oxidative stress in aging and healthspan. Longev Healthspan. 3:6.
  • Dai DF, Chiao YA, Martin GM, Marcinek DJ, Basisty N, Quarles EK, Rabinovitch PS. 2017. Mitochondrial-targeted catalase: extended longevity and the roles in various disease models. Prog Mol Biol Transl Sci. 146:203–241.
  • Dai DF, Hsieh EJ, Chen T, Menendez LG, Basisty NB, Tsai L, Beyer RP, Crispin DA, Shulman NJ, Szeto HH, et al. 2013. Global proteomics and pathway analysis of pressure-overload-induced heart failure and its attenuation by mitochondrial-targeted peptides. Circ Heart Fail. 6(5):1067–1076.
  • Dai DF, Hsieh EJ, Liu Y, Chen T, Beyer RP, Chin MT, MacCoss MJ, Rabinovitch PS. 2012. Mitochondrial proteome remodelling in pressure overload-induced heart failure: the role of mitochondrial oxidative stress. Cardiovasc Res. 93(1):79–88.
  • Dai DF, Johnson SC, Villarin JJ, Chin MT, Nieves-Cintron M, Chen T, Marcinek DJ, Dorn GW, 2nd Kang YJ, Prolla TA, et al. 2011. Mitochondrial oxidative stress mediates angiotensin II-induced cardiac hypertrophy and Galphaq overexpression-induced heart failure. Circ Res. 108(7):837–846.
  • Dai DF, Rabinovitch P. 2011. Mitochondrial oxidative stress mediates induction of autophagy and hypertrophy in angiotensin-II treated mouse hearts. Autophagy. 7(8):917–918.
  • Dai DF, Santana LF, Vermulst M, Tomazela DM, Emond MJ, MacCoss MJ, Gollahon K, Martin GM, Loeb LA, Ladiges WC, et al. 2009. Overexpression of catalase targeted to mitochondria attenuates murine cardiac aging. Circulation. 119(21):2789–2797.
  • Dai W, Shi J, Gupta RC, Sabbah HN, Hale SL, Kloner RA. 2014. Bendavia, a mitochondria-targeting peptide, improves postinfarction cardiac function, prevents adverse left ventricular remodeling, and restores mitochondria-related gene expression in rats. J Cardiovasc Pharmacol. 64(6):543–553.
  • Dallas ML, Scragg JL, Peers C. 2009. Inhibition of L-type Ca(2+) channels by carbon monoxide. Adv Exp Med Biol. 648:89–95.
  • Dare AJ, Bolton EA, Pettigrew GJ, Bradley JA, Saeb-Parsy K, Murphy MP. 2015. Protection against renal ischemia-reperfusion injury in vivo by the mitochondria targeted antioxidant MitoQ. Redox Biol. 5:163–168.
  • Dare AJ, Logan A, Prime TA, Rogatti S, Goddard M, Bolton EM, Bradley JA, Pettigrew GJ, Murphy MP, Saeb-Parsy K. 2015. The mitochondria-targeted anti-oxidant MitoQ decreases ischemia-reperfusion injury in a murine syngeneic heart transplant model. J Heart Lung Transplant. 34(11):1471–1480.
  • Dashdorj A, Jyothi KR, Lim S, Jo A, Nguyen MN, Ha J, Yoon KS, Kim HJ, Park JH, Murphy MP, et al. 2013. Mitochondria-targeted antioxidant MitoQ ameliorates experimental mouse colitis by suppressing NLRP3 inflammasome-mediated inflammatory cytokines. BMC Med. 11(1):178.
  • Daubert MA, Yow E, Dunn G, Marchev S, Barnhart H, Douglas PS, O’Connor C, Goldstein S, Udelson JE, Sabbah HN. 2017. Novel mitochondria-targeting peptide in heart failure treatment: a randomized, placebo-controlled trial of Elamipretide. Circ Heart Fail. 10(12)
  • Davis CA, Nick HS, Agarwal A. 2001. Manganese superoxide dismutase attenuates cisplatin-induced renal injury: importance of superoxide. J Am Soc Nephrol. 12(12):2683–2690.
  • De Benedictis G, Carotenuto L, Carrieri G, De Luca M, Falcone E, Rose G, Cavalcanti S, Corsonello F, Feraco E, Baggio G, et al. 1998. Gene/longevity association studies at four autosomal loci (REN, THO, PARP, SOD2). Eur J Hum Genet. 6(6):534–541.
  • de Mendonca E, Salazar Alcala E, Fernandez-Mestre M. 2016. Role of genes GSTM1, GSTT1, and MnSOD in the development of late-onset Alzheimer disease and their relationship with APOE*4. Neurologia. 31(8):535–542.
  • De Simoni S, Goemaere J, Knoops B. 2008. Silencing of peroxiredoxin 3 and peroxiredoxin 5 reveals the role of mitochondrial peroxiredoxins in the protection of human neuroblastoma SH-SY5Y cells toward MPP+. Neurosci Lett. 433(3):219–224.
  • Delco ML, Bonnevie ED, Szeto HS, Bonassar LJ, Fortier LA. 2018. Mitoprotective therapy preserves chondrocyte viability and prevents cartilage degeneration in an ex vivo model of posttraumatic osteoarthritis. J Orthop Res. 36(8):2147–2156.
  • Demianenko IA, Vasilieva TV, Domnina LV, Dugina VB, Egorov MV, Ivanova OY, Ilinskaya OP, Pletjushkina OY, Popova EN, Sakharov IY, et al. 2010. Novel mitochondria-targeted antioxidants, “Skulachev-ion” derivatives, accelerate dermal wound healing in animals. Biochemistry Moscow. 75(3):274–280.
  • Demyanenko IA, Popova EN, Zakharova VV, Ilyinskaya OP, Vasilieva TV, Romashchenko VP, Fedorov AV, Manskikh VN, Skulachev MV, Zinovkin RA, et al. 2015. Mitochondria-targeted antioxidant SkQ1 improves impaired dermal wound healing in old mice. Aging (Albany NY). 7(7):475–485.
  • Demyanenko IA, Zakharova VV, Ilyinskaya OP, Vasilieva TV, Fedorov AV, Manskikh VN, Zinovkin RA, Pletjushkina OY, Chernyak BV, Skulachev VP, et al. 2017. Mitochondria-targeted antioxidant SkQ1 improves dermal wound healing in genetically diabetic mice. Oxid Med Cell Longev. 2017:6408278.
  • Deng FY, Lei SF, Chen XD, Tan LJ, Zhu XZ, Deng HW. 2011. An integrative study ascertained SOD2 as a susceptibility gene for osteoporosis in Chinese. J Bone Miner Res. 26(11):2695–2701.
  • Despotovic M, Stoimenov TJ, Stankovic I, Pavlovic D, Sokolovic D, Cvetkovic T, Kocic G, Basic J, Veljkovic A, Djordjevic B. 2015. Gene polymorphisms of tumor necrosis factor alpha and antioxidant enzymes in bronchial asthma. Adv Clin Exp Med. 24(2):251–256.
  • Detaille D, Pasdois P, Semont A, Dos Santos P, Diolez P. 2019. An old medicine as a new drug to prevent mitochondrial complex I from producing oxygen radicals. PLoS One. 14(5):e0216385.
  • Detaille D, Vial G, Borel AL, Cottet-Rouselle C, Hallakou-Bozec S, Bolze S, Fouqueray P, Fontaine E. 2016. Imeglimin prevents human endothelial cell death by inhibiting mitochondrial permeability transition without inhibiting mitochondrial respiration. Cell Death Discov. 2:15072.
  • Dhiman M, Wan X, Popov VL, Vargas G, Garg NJ. 2013. MnSODtg mice control myocardial inflammatory and oxidative stress and remodeling responses elicited in chronic Chagas disease. J Am Heart Assoc. 2(5):e000302.
  • Diebold L, Chandel NS. 2016. Mitochondrial ROS regulation of proliferating cells. Free Radic Biol Med. 100:86–93.
  • Dikalov SI, Nazarewicz RR. 2013. Angiotensin II-induced production of mitochondrial reactive oxygen species: potential mechanisms and relevance for cardiovascular disease. Antioxid Redox Signaling. 19(10):1085–1094.
  • Dikalov SI, Nazarewicz RR, Bikineyeva A, Hilenski L, Lassegue B, Griendling KK, Harrison DG, Dikalova AE. 2014. Nox2-induced production of mitochondrial superoxide in angiotensin II-mediated endothelial oxidative stress and hypertension. Antioxid Redox Signal. 20(2):281–294.
  • Dikalova AE, Bikineyeva AT, Budzyn K, Nazarewicz RR, McCann L, Lewis W, Harrison DG, Dikalov SI. 2010. Therapeutic targeting of mitochondrial superoxide in hypertension. Circ Res. 107(1):106–116.
  • Dikalova AE, Itani HA, Nazarewicz RR, McMaster WG, Flynn CR, Uzhachenko R, Fessel JP, Gamboa JL, Harrison DG, Dikalov SI. 2017. Sirt3 impairment and SOD2 hyperacetylation in vascular oxidative stress and hypertension. Circ Res. 121(5):564–574.
  • Dikalova AE, Kirilyuk IA, Dikalov SI. 2015. Antihypertensive effect of mitochondria-targeted proxyl nitroxides. Redox Biol. 4:355–362.
  • Ding W, Guo H, Xu C, Wang B, Zhang M, Ding F. 2016. Mitochondrial reactive oxygen species-mediated NLRP3 inflammasome activation contributes to aldosterone-induced renal tubular cells injury. Oncotarget. 7(14):17479–17491.
  • Ding Y, Jiang Z, Xia B, Zhang L, Zhang C, Leng J. 2019. Mitochondria-targeted antioxidant therapy for an animal model of PCOS-IR. Int J Mol Med. 43(1):316–324.
  • Ding W, Liu T, Bi X, Zhang Z. 2017. Mitochondria-targeted antioxidant Mito-Tempo protects against aldosterone-induced renal injury in vivo. Cell Physiol Biochem. 44(2):741–750.
  • Diolez P, Detaille D, Marin F, Petitjean O, inventors. 2019 Sep. 12, 2019. Compounds for the treatment of diseases linked to mitochondrial reactive oxygen species (ROS) production. United States Patent application US 2019/0274997 A1 patent US 2019/0274997 A1.
  • Dirain CO, Ng M, Milne-Davies B, Joseph JK, Antonelli PJ. 2018. Evaluation of mitoquinone for protecting against amikacin-induced ototoxicity in guinea pigs. Otol Neurotol. 39(1):111–118.
  • Doctrow SR, Liesa M, Melov S, Shirihai OS, Tofilon P. 2012. Salen Mn complexes are superoxide dismutase/catalase mimetics that protect the mitochondria. Curr Inorg Chem. 2(3):325–334.
  • Dolgachev V, Oberley LW, Huang TT, Kraniak JM, Tainsky MA, Hanada K, Separovic D. 2005. A role for manganese superoxide dismutase in apoptosis after photosensitization. Biochem Biophys Res Commun. 332(2):411–417.
  • Domazetovic V, Fontani F, Marcucci G, Iantomasi T, Brandi ML, Vincenzini MT. 2017. Estrogen inhibits starvation-induced apoptosis in osteocytes by a redox-independent process involving association of JNK and glutathione S-transferase P1-1. FEBS Open Bio. 7(5):705–718.
  • Drose S, Brandt U. 2012. Molecular mechanisms of superoxide production by the mitochondrial respiratory chain. Adv Exp Med Biol. 748:145–169.
  • Du XL, Edelstein D, Dimmeler S, Ju Q, Sui C, Brownlee M. 2001. Hyperglycemia inhibits endothelial nitric oxide synthase activity by posttranslational modification at the Akt site. J Clin Invest. 108(9):1341–1348.
  • Du XL, Edelstein D, Rossetti L, Fantus IG, Goldberg H, Ziyadeh F, Wu J, Brownlee M. 2000. Hyperglycemia-induced mitochondrial superoxide overproduction activates the hexosamine pathway and induces plasminogen activator inhibitor-1 expression by increasing Sp1 glycosylation. Proc Natl Acad Sci USA. 97(22):12222–12226.
  • Du K, Farhood A, Jaeschke H. 2017. Mitochondria-targeted antioxidant Mito-Tempo protects against acetaminophen hepatotoxicity. Arch Toxicol. 91(2):761–773.
  • Du K, Ramachandran A, Jaeschke H. 2016. Oxidative stress during acetaminophen hepatotoxicity: Sources, pathophysiological role and therapeutic potential. Redox Biol. 10:148–156.
  • Du K, Ramachandran A, Weemhoff JL, Woolbright BL, Jaeschke AH, Chao X, Ding WX, Jaeschke H. 2019. Mito-tempo protects against acute liver injury but induces limited secondary apoptosis during the late phase of acetaminophen hepatotoxicity. Arch Toxicol. 93(1):163–178.
  • Duan SB, Yang SK, Zhou QY, Pan P, Zhang H, Liu F, Xu XQ. 2013. Mitochondria-targeted peptides prevent on contrast-induced acute kidney injury in the rats with hypercholesterolemia. Ren Fail. 35(8):1124–1129.
  • Ducasa GM, Mitrofanova A, Mallela SK, Liu X, Molina J, Sloan A, Pedigo CE, Ge M, Santos JV, Hernandez Y, et al. 2019. ATP-binding cassette A1 deficiency causes cardiolipin-driven mitochondrial dysfunction in podocytes. J Clin Invest. 129(8):3387–3400.
  • Dumont M, Wille E, Stack C, Calingasan NY, Beal MF, Lin MT. 2009. Reduction of oxidative stress, amyloid deposition, and memory deficit by manganese superoxide dismutase overexpression in a transgenic mouse model of Alzheimer’s disease. FASEB J. 23(8):2459–2466.
  • Echtay KS, Murphy MP, Smith RA, Talbot DA, Brand MD. 2002. Superoxide activates mitochondrial uncoupling protein 2 from the matrix side. Studies using targeted antioxidants. J Biol Chem. 277(49):47129–47135.
  • Echtay KS, Roussel D, St-Pierre J, Jekabsons MB, Cadenas S, Stuart JA, Harper JA, Roebuck SJ, Morrison A, Pickering S, et al. 2002. Superoxide activates mitochondrial uncoupling proteins. Nature. 415(6867):96–99.
  • Egan KM, Thompson PA, Titus-Ernstoff L, Moore JH, Ambrosone CB. 2003. MnSOD polymorphism and breast cancer in a population-based case-control study. Cancer Lett. 199(1):27–33.
  • Eirin A, Ebrahimi B, Zhang X, Zhu X-Y, Woollard JR, He Q, Textor SC, Lerman A, Lerman LO. 2014. Mitochondrial protection restores renal function in swine atherosclerotic renovascular disease. Cardiovasc Res. 103(4):461–472.
  • Eirin A, Li Z, Zhang X, Krier JD, Woollard JR, Zhu X-Y, Tang H, Herrmann SM, Lerman A, Textor SC, et al. 2012. A mitochondrial permeability transition pore inhibitor improves renal outcomes after revascularization in experimental atherosclerotic renal artery stenosis. Hypertension. 60(5):1242–1249.
  • Eismann T, Huber N, Shin T, Kuboki S, Galloway E, Wyder M, Edwards MJ, Greis KD, Shertzer HG, Fisher AB, et al. 2009. Peroxiredoxin-6 protects against mitochondrial dysfunction and liver injury during ischemia-reperfusion in mice. Am J Physiol Gastrointest Liver Physiol. 296(2):G266–274.
  • el-Masry TM, Zahra MA, el-Tawil MM, Khalifa RA. 2005. Manganese superoxide dismutase alanine to valine polymorphism and risk of neuropathy and nephropathy in Egyptian type 1 diabetic patients. Rev Diabet Stud. 2(2):70–74.
  • Elnakish MT, Schultz EJ, Gearinger RL, Saad NS, Rastogi N, Ahmed AA, Mohler PJ, Janssen PM. 2015. Differential involvement of various sources of reactive oxygen species in thyroxin-induced hemodynamic changes and contractile dysfunction of the heart and diaphragm muscles. Free Radic Biol Med. 83:252–261.
  • Elsakka NE, Webster NR, Galley HF. 2007. Polymorphism in the manganese superoxide dismutase gene. Free Radic Res. 41(7):770–778.
  • Epperly MW, Bernarding M, Gretton J, Jefferson M, Nie S, Greenberger JS. 2003. Overexpression of the transgene for manganese superoxide dismutase (MnSOD) in 32D cl 3 cells prevents apoptosis induction by TNF-alpha, IL-3 withdrawal, and ionizing radiation. Exp Hematol. 31(6):465–474.
  • Epperly MW, Bray JA, Krager S, Berry LM, Gooding W, Engelhardt JF, Zwacka R, Travis EL, Greenberger JS. 1999. Intratracheal injection of adenovirus containing the human MnSOD transgene protects athymic nude mice from irradiation-induced organizing alveolitis. Int J Radiat Oncol Biol Phys. 43(1):169–181.
  • Epperly MW, Defilippi S, Sikora C, Gretton J, Greenberger JS. 2002. Radioprotection of lung and esophagus by overexpression of the human manganese superoxide dismutase transgene. Mil Med. 167(2 Suppl):71–73.
  • Epperly MW, Epstein CJ, Travis EL, Greenberger JS. 2000. Decreased pulmonary radiation resistance of manganese superoxide dismutase (MnSOD)-deficient mice is corrected by human manganese superoxide dismutase-Plasmid/Liposome (SOD2-PL) intratracheal gene therapy. Radiat Res. 154(4):365–374.
  • Epperly MW, Melendez JA, Zhang X, Nie S, Pearce L, Peterson J, Franicola D, Dixon T, Greenberger BA, Komanduri P, et al. 2009. Mitochondrial targeting of a catalase transgene product by plasmid liposomes increases radioresistance in vitro and in vivo. Radiat Res. 171(5):588–595.
  • Epstein JB, Decoteau WE, Wilkinson A. 1983. Effect of Sialor in treatment of xerostomia in Sjogren’s syndrome. Oral Surg Oral Med Oral Pathol. 56(5):495–499.
  • Ergen HA, Narter F, Timirci O, Isbir T. 2007. Effects of manganase superoxide dismutase Ala-9Val polymorphism on prostate cancer: a case-control study. Anticancer Res. 27(2):1227–1230.
  • Escobales N, Nunez RE, Jang S, Parodi-Rullan R, Ayala-Pena S, Sacher JR, Skoda EM, Wipf P, Frontera W, Javadov S. 2014. Mitochondria-targeted ROS scavenger improves post-ischemic recovery of cardiac function and attenuates mitochondrial abnormalities in aged rats. J Mol Cell Cardiol. 77:136–146.
  • Escribano-Lopez I, Banuls C, Diaz-Morales N, Iannantuoni F, Rovira-Llopis S, Gomis R, Rocha M, Hernandez-Mijares A, Murphy MP, Victor VM. 2019. The mitochondria-targeted antioxidant mitoQ modulates mitochondrial function and endoplasmic reticulum stress in pancreatic β Cells Exposed to Hyperglycaemia. Cell Physiol Biochem. 52(2):186–197.
  • Escribano-Lopez I, Diaz-Morales N, Iannantuoni F, Lopez-Domenech S, de Maranon AM, Abad-Jimenez Z, Banuls C, Rovira-Llopis S, Herance JR, Rocha M, et al. 2018. The mitochondrial antioxidant SS-31 increases SIRT1 levels and ameliorates inflammation, oxidative stress and leukocyte-endothelium interactions in type 2 diabetes. Sci Rep. 8(1):15862.
  • Escribano-Lopez I, Diaz-Morales N, Rovira-Llopis S, de Maranon AM, Orden S, Alvarez A, Banuls C, Rocha M, Murphy MP, Hernandez-Mijares A, et al. 2016. The mitochondria-targeted antioxidant MitoQ modulates oxidative stress, inflammation and leukocyte-endothelium interactions in leukocytes isolated from type 2 diabetic patients. Redox Biol. 10:200–205.
  • Esih K, Goricar K, Dolzan V, Rener-Primec Z. 2017. Antioxidant polymorphisms do not influence the risk of epilepsy or its drug resistance after neonatal hypoxic-ischemic brain injury. Seizure. 46:38–42.
  • Esposito L, Raber J, Kekonius L, Yan F, Yu GQ, Bien-Ly N, Puolivali J, Scearce LK, Masliah E, Mucke L. 2006. Reduction in mitochondrial superoxide dismutase modulates Alzheimer’s disease-like pathology and accelerates the onset of behavioral changes in human amyloid precursor protein transgenic mice. J Neurosci. 26(19):5167–5179.
  • Esterberg R, Linbo T, Pickett SB, Wu P, Ou HC, Rubel EW, Raible DW. 2016. Mitochondrial calcium uptake underlies ROS generation during aminoglycoside-induced hair cell death. J Clin Invest. 126(9):3556–3566.
  • Esworthy RS, Ho YS, Chu FF. 1997. The Gpx1 gene encodes mitochondrial glutathione peroxidase in the mouse liver. Arch Biochem Biophys. 340(1):59–63.
  • Ezzikouri S, El Feydi AE, Chafik A, Afifi R, El Kihal L, Benazzouz M, Hassar M, Pineau P, Benjelloun S. 2008. Genetic polymorphism in the manganese superoxide dismutase gene is associated with an increased risk for hepatocellular carcinoma in HCV-infected Moroccan patients. Mutat Res. 649(1–2):1–6.
  • Fang L, Bai C, Chen Y, Dai J, Xiang Y, Ji X, Huang C, Dong Q. 2014. Inhibition of ROS production through mitochondria-targeted antioxidant and mitochondrial uncoupling increases post-thaw sperm viability in yellow catfish. Cryobiology. 69(3):386–393.
  • Fang D, Heslop K, Morris M, DeHart D, Monika Beck Gooz MB, Lemasters JJ, Maldonado EN. 2017. Oxidative stress induced by VDAC opening in cancer cells depends on cytosolic free tubulin and is blocked by ROS scavenging and suppression of superoxide formation by Complex III [Abstract]. Biophys J. 112(3):324A.
  • Fang X, Wang H, Han D, Xie E, Yang X, Wei J, Gu S, Gao F, Zhu N, Yin X, et al. 2019. Ferroptosis as a target for protection against cardiomyopathy. Proc Natl Acad Sci USA. 116(7):2672–2680.
  • Fang J, Wong HS, Brand MD. 2020. Production of superoxide and hydrogen peroxide in the mitochondrial matrix is dominated by site IQ of complex I in diverse cell lines. Redox Biol. 37:101722.
  • Farin FM, Hitosis Y, Hallagan SE, Kushleika J, Woods JS, Janssen PS, Smith-Weller T, Franklin GM, Swanson PD, Checkoway H. 2001. Genetic polymorphisms of superoxide dismutase in Parkinson’s disease. Mov Disord. 16(4):705–707.
  • Farnaghi S, Prasadam I, Cai G, Friis T, Du Z, Crawford R, Mao X, Xiao Y. 2017. Protective effects of mitochondria-targeted antioxidants and statins on cholesterol-induced osteoarthritis. FASEB J. 31(1):356–367.
  • Fatemie S, Goh J, Pettan-Brewer C, Ladiges W. 2012. Breast tumors in PyMT transgenic mice expressing mitochondrial catalase have decreased labeling for macrophages and endothelial cells. Pathobiol Aging Age Relat Dis. 2. doi:10.3402/pba.v2i0.17391
  • Faure C, Leveille P, Dupont C, Julia C, Chavatte-Palmer P, Alifert G, Sutton A, Levy R, Alifert Group 2014. Are superoxide dismutase 2 and nitric oxide synthase polymorphisms associated with idiopathic infertility? Antioxid Redox Signal. 21(4):565–569.
  • Feillet-Coudray C, Fouret G, Ebabe Elle R, Rieusset J, Bonafos B, Chabi B, Crouzier D, Zarkovic K, Zarkovic N, Ramos J, et al. 2014. The mitochondrial-targeted antioxidant MitoQ ameliorates metabolic syndrome features in obesogenic diet-fed rats better than Apocynin or Allopurinol. Free Radic Res. 48(10):1232–1246.
  • Fennell JP, Brosnan MJ, Frater AJ, Hamilton CA, Alexander MY, Nicklin SA, Heistad DD, Baker AH, Dominiczak AF. 2002. Adenovirus-mediated overexpression of extracellular superoxide dismutase improves endothelial dysfunction in a rat model of hypertension. Gene Ther. 9(2):110–117.
  • Filograna R, Godena VK, Sanchez-Martinez A, Ferrari E, Casella L, Beltramini M, Bubacco L, Whitworth AJ, Bisaglia M. 2016. Superoxide dismutase (SOD)-mimetic M40403 Is protective in cell and fly models of paraquat toxicity: implications for Parkinson disease. J Biol Chem. 291(17):9257–9267.
  • Fink BD, Guo DF, Kulkarni CA, Rahmouni K, Kerns RJ, Sivitz WI. 2017. Metabolic effects of a mitochondrial-targeted coenzyme Q analog in high fat fed obese mice. Pharmacol Res Perspect. 5(2):e00301.
  • Fink BD, Herlein JA, Guo DF, Kulkarni C, Weidemann BJ, Yu L, Grobe JL, Rahmouni K, Kerns RJ, Sivitz WI. 2014. A mitochondrial-targeted coenzyme Q analog prevents weight gain and ameliorates hepatic dysfunction in high-fat-fed mice. J Pharmacol Exp Ther. 351(3):699–708.
  • Fink MP, Macias CA, Xiao J, Tyurina YY, Jiang J, Belikova N, Delude RL, Greenberger JS, Kagan VE, Wipf P. 2007. Hemigramicidin-TEMPO conjugates: novel mitochondria-targeted anti-oxidants. Biochem Pharmacol. 74(6):801–809.
  • Fock E, Bachteeva V, Lavrova E, Parnova R. 2018. Mitochondrial-targeted antioxidant MitoQ prevents E. coli lipopolysaccharide-induced accumulation of triacylglycerol and lipid droplets biogenesis in epithelial cells. J Lipids. 2018:5745790.
  • Fong CS, Wu RM, Shieh JC, Chao YT, Fu YP, Kuao CL, Cheng CW. 2007. Pesticide exposure on southwestern Taiwanese with MnSOD and NQO1 polymorphisms is associated with increased risk of Parkinson’s disease. Clin Chim Acta. 378(1-2):136–141.
  • Fontaine E, Detaille D, Vial G. 2016. Preventing cell death with a ‘check valve’ in mitochondrial complex I? Cell Death Dis. 7:e2165.
  • Formentini L, Santacatterina F, Nunez de Arenas C, Stamatakis K, Lopez-Martinez D, Logan A, Fresno M, Smits R, Murphy MP, Cuezva JM. 2017. Mitochondrial ROS production protects the intestine from inflammation through functional M2 macrophage polarization. Cell Rep. 19(6):1202–1213.
  • Fortner KA, Blanco LP, Buskiewicz I, Huang N, Gibson PC, Cook DL, Pedersen HL, Yuen PST, Murphy MP, Perl A, et al. 2020. Targeting mitochondrial oxidative stress with MitoQ reduces NET formation and kidney disease in lupus-prone MRL-lpr mice. Lupus Sci Med. 7(1):e000387.
  • Fortunato G, Marciano E, Zarrilli F, Mazzaccara C, Intrieri M, Calcagno G, Vitale DF, La Manna P, Saulino C, Marcelli V, et al. 2004. Paraoxonase and superoxide dismutase gene polymorphisms and noise-induced hearing loss. Clin Chem. 50(11):2012–2018.
  • Fouqueray P, Leverve X, Fontaine E, Baquié M, Wollheim C, Lebovitz H, Bozec S. 2011. Imeglimin - a new oral anti-diabetic that targets the three key defects of type 2 diabetes. J Diabetes Metab. 02(04):126–133.
  • Fouret G, Tolika E, Lecomte J, Bonafos B, Aoun M, Murphy MP, Ferreri C, Chatgilialoglu C, Dubreucq E, Coudray C, et al. 2015. The mitochondrial-targeted antioxidant, MitoQ, increases liver mitochondrial cardiolipin content in obesogenic diet-fed rats. Biochim Biophys Acta. 1847(10):1025–1035.
  • Fu C, Hao S, Liu Z, Xie L, Wu X, Wu X, Li S. 2020. SOD2 ameliorates pulmonary hypertension in a murine model of sleep apnea via suppressing expression of NLRP3 in CD11b(+) cells. Respir Res. 21(1):9.
  • Fujimoto K, Kumagai K, Ito K, Arakawa S, Ando Y, Oda S, Yamoto T, Manabe S. 2009. Sensitivity of liver injury in heterozygous Sod2 knockout mice treated with troglitazone or acetaminophen. Toxicol Pathol. 37(2):193–200.
  • Fujimoto C, Yamasoba T. 2019. Mitochondria-targeted antioxidants for treatment of hearing loss: a systematic review. Antioxidants (Basel). 8(4):109.
  • Fujimura M, Morita-Fujimura Y, Kawase M, Copin JC, Calagui B, Epstein CJ, Chan PH. 1999. Manganese superoxide dismutase mediates the early release of mitochondrial cytochrome C and subsequent DNA fragmentation after permanent focal cerebral ischemia in mice. J Neurosci. 19(9):3414–3422.
  • Fukui M, Zhu BT. 2010. Mitochondrial superoxide dismutase SOD2, but not cytosolic SOD1, plays a critical role in protection against glutamate-induced oxidative stress and cell death in HT22 neuronal cells. Free Rad Biol Med. 48(6):821–830.
  • Galecki P, Pietras T, Szemraj J. 2006. Manganese superoxide dismutase gene (MnSOD) polimorphism in schizophrenics with tardive dyskinesia from central Poland. Psychiatr Pol. 40(5):937–948.
  • Gałecki P, Smigielski J, Florkowski A, Bobińska K, Pietras T, Szemraj J. 2010. Analysis of two polymorphisms of the manganese superoxide dismutase gene (Ile-58Thr and Ala-9Val) in patients with recurrent depressive disorder. Psychiatry Res. 179(1):43–46.
  • Gallagher CJ, Ahn K, Knipe AL, Dyer AM, Richie JP, Jr., Lazarus P, Muscat JE. 2009. Association between haplotypes of manganese superoxide dismutase (SOD2), smoking, and lung cancer risk. Free Rad Biol Med. 46(1):20–24.
  • Gamarra D, Elcoroaristizabal X, Fernandez-Martinez M, de Pancorbo MM. 2015. Association of the C47T polymorphism in SOD2 with amnestic mild cognitive impairment and Alzheimer’s disease in carriers of the APOEepsilon4 allele. Dis Markers. 2015:746329.
  • Gan L, Wang Z, Si J, Zhou R, Sun C, Liu Y, Ye Y, Zhang Y, Liu Z, Zhang H. 2018. Protective effect of mitochondrial-targeted antioxidant MitoQ against iron ion (56)Fe radiation induced brain injury in mice. Toxicol Appl Pharmacol. 341:1–7.
  • Gane EJ, Weilert F, Orr DW, Keogh GF, Gibson M, Lockhart MM, Frampton CM, Taylor KM, Smith RA, Murphy MP. 2010. The mitochondria-targeted anti-oxidant mitoquinone decreases liver damage in a phase II study of hepatitis C patients. Liver Int. 30(7):1019–1026.
  • Ganguly E, Aljunaidy MM, Kirschenman R, Spaans F, Morton JS, Phillips TEJ, Case CP, Cooke CM, Davidge ST. 2019. Sex-specific effects of nanoparticle-encapsulated MitoQ (nMitoQ) delivery to the placenta in a rat model of fetal hypoxia. Front Physiol. 10:562.
  • Ganini D, Santos JH, Bonini MG, Mason RP. 2018. Switch of mitochondrial superoxide dismutase into a prooxidant peroxidase in manganese-deficient cells and mice. Cell Chem Biol. 25(4):413–425.
  • Gao J, Feng Z, Wang X, Zeng M, Liu J, Han S, Xu J, Chen L, Cao K, Long J, et al. 2018. SIRT3/SOD2 maintains osteoblast differentiation and bone formation by regulating mitochondrial stress. Cell Death Differ. 25(2):229–240.
  • Gao W, Tan J, Hüls A, Ding A, Liu Y, Matsui MS, Vierkötter A, Krutmann J, Schikowski T, Jin L, et al. 2017. Genetic variants associated with skin aging in the Chinese Han population. J Dermatol Sci. 86(1):21–29.
  • Gardner PR. 2002. Aconitase: sensitive target and measure of superoxide. Meth Enzymol. 349:9–23.
  • Gardner R, Salvador A, Moradas-Ferreira P. 2002. Why does SOD overexpression sometimes enhance, sometimes decrease, hydrogen peroxide production? A minimalist explanation. Free Rad Biol Med. 32(12):1351–1357.
  • Gaudet MM, Gammon MD, Santella RM, Britton JA, Teitelbaum SL, Eng SM, Terry MB, Bensen JT, Schroeder J, Olshan AF, et al. 2005. MnSOD Val-9Ala genotype, pro- and anti-oxidant environmental modifiers, and breast cancer among women on Long Island, New York. Cancer Causes Control. 16(10):1225–1234.
  • Genrikhs EE, Stelmashook EV, Alexandrova OP, Novikova SV, Voronkov DN, Glibka YA, Skulachev VP, Isaev NK. 2019. The single intravenous administration of mitochondria-targeted antioxidant SkQR1 after traumatic brain injury attenuates neurological deficit in rats. Brain Res Bull. 148:100–108.
  • Gentschew L, Flachsbart F, Kleindorp R, Badarinarayan N, Schreiber S, Nebel A. 2013. Polymorphisms in the superoxidase dismutase genes reveal no association with human longevity in Germans: a case-control association study. Biogerontology. 14(6):719–727.
  • Gero D, Szoleczky P, Suzuki K, Modis K, Olah G, Coletta C, Szabo C. 2013. Cell-based screening identifies paroxetine as an inhibitor of diabetic endothelial dysfunction. Diabetes. 62(3):953–964.
  • Gerő D, Torregrossa R, Perry A, Waters A, Le-Trionnaire S, Whatmore JL, Wood M, Whiteman M. 2016. The novel mitochondria-targeted hydrogen sulfide (H2S) donors AP123 and AP39 protect against hyperglycemic injury in microvascular endothelial cells in vitro. Pharmacol Res. 113(Pt A):186–198.
  • Ghosh A, Chandran K, Kalivendi SV, Joseph J, Antholine WE, Hillard CJ, Kanthasamy A, Kanthasamy A, Kalyanaraman B. 2010. Neuroprotection by a mitochondria-targeted drug in a Parkinson’s disease model. Free Rad Biol Med. 49(11):1674–1684.
  • Gibson CM, Giugliano RP, Kloner RA, Bode C, Tendera M, Jánosi A, Merkely B, Godlewski J, Halaby R, Korjian S, et al. 2016. EMBRACE STEMI study: a Phase 2a trial to evaluate the safety, tolerability, and efficacy of intravenous MTP-131 on reperfusion injury in patients undergoing primary percutaneous coronary intervention. Eur Heart J. 37(16):1296–1303.
  • Gilliam LA, Lark DS, Reese LR, Torres MJ, Ryan TE, Lin CT, Cathey BL, Neufer PD. 2016. Targeted overexpression of mitochondrial catalase protects against cancer chemotherapy-induced skeletal muscle dysfunction. Am J Physiol Endocrinol Metab. 311(2):E293–E301.
  • Gilliam LA, Moylan JS, Patterson EW, Smith JD, Wilson AS, Rabbani Z, Reid MB. 2012. Doxorubicin acts via mitochondrial ROS to stimulate catabolism in C2C12 myotubes. Am J Physiol, Cell Physiol. 302(1):C195–C202.
  • Gioscia-Ryan RA, Battson ML, Cuevas LM, Eng JS, Murphy MP, Seals DR. 2018. Mitochondria-targeted antioxidant therapy with MitoQ ameliorates aortic stiffening in old mice. J Appl Physiol. 124(5):1194–1202.
  • Gioscia-Ryan RA, LaRocca TJ, Sindler AL, Zigler MC, Murphy MP, Seals DR. 2014. Mitochondria-targeted antioxidant (MitoQ) ameliorates age-related arterial endothelial dysfunction in mice. J Physiol. 592(12):2549–2561.
  • Gitik M, Srivastava V, Hodgkinson CA, Shen PH, Goldman D, Meyerhoff DJ. 2016. Association of superoxide dismutase 2 (SOD2) genotype with gray matter volume shrinkage in chronic alcohol users: replication and further evaluation of an addiction gene panel. Int J Neuropsychopharmacol. 19(9):pyw033.
  • Giusti B, Vestrini A, Poggi C, Magi A, Pasquini E, Abbate R, Dani C. 2012. Genetic polymorphisms of antioxidant enzymes as risk factors for oxidative stress-associated complications in preterm infants. Free Radic Res. 46(9):1130–1139.
  • Glenert U. 1992. Effects of chronic anethole trithione and amitriptyline treatment on rat parotid gland signalling. Eur J Pharmacol. 226(1):43–52.
  • Glover M, Hebert VY, Nichols K, Xue SY, Thibeaux TM, Zavecz JA, Dugas TR. 2014. Overexpression of mitochondrial antioxidant manganese superoxide dismutase (MnSOD) provides protection against AZT- or 3TC-induced endothelial dysfunction. Antiviral Res. 111:136–142.
  • Goh J, Enns L, Fatemie S, Hopkins H, Morton J, Pettan-Brewer C, Ladiges W. 2011. Mitochondrial targeted catalase suppresses invasive breast cancer in mice. BMC Cancer. 11:191.
  • Goh KY, He L, Song J, Jinno M, Rogers AJ, Sethu P, Halade GV, Rajasekaran NS, Liu X, Prabhu SD, et al. 2019. Mitoquinone ameliorates pressure overload-induced cardiac fibrosis and left ventricular dysfunction in mice. Redox Biol. 21:101100.
  • Goncalves RL, Quinlan CL, Perevoshchikova IV, Hey-Mogensen M, Brand MD. 2015. Sites of superoxide and hydrogen peroxide production by muscle mitochondria assessed ex vivo under conditions mimicking rest and exercise. J Biol Chem. 290(1):209–227.
  • Goncalves RLS, Watson MA, Wong HS, Orr AL, Brand MD. 2020. The use of site-specific suppressors to measure the relative contributions of different mitochondrial sites to skeletal muscle superoxide and hydrogen peroxide production. Redox Biol. 28:101341.
  • Goodman JH. 2016. Neuroprotection and anti-epileptogenesis with mitochondria-targeted antioxidant. Defense Technical Information Center Technical Report https://appsdticmil/dtic/tr/fulltext/u2/1020626pdf.
  • Goto H, Nishikawa T, Sonoda K, Kondo T, Kukidome D, Fujisawa K, Yamashiro T, Motoshima H, Matsumura T, Tsuruzoe K, et al. 2008. Endothelial MnSOD overexpression prevents retinal VEGF expression in diabetic mice. Biochem Biophys Res Commun. 366(3):814–820.
  • Gottlieb MG, Schwanke CH, Santos AF, Jobim PF, Mussel DP, da Cruz IB. 2005. Association among oxidized LDL levels, MnSOD, apolipoprotein E polymorphisms, and cardiovascular risk factors in a south Brazilian region population. Genet Mol Res. 4(4):691–703.
  • Gouni-Berthold I, Berthold HK, Huh JY, Berman R, Spenrath N, Krone W, Mantzoros CS. 2013. Effects of lipid-lowering drugs on irisin in human subjects in vivo and in human skeletal muscle cells ex vivo. PLoS One. 8(9):e72858.
  • Graham D, Huynh NN, Hamilton CA, Beattie E, Smith RA, Cocheme HM, Murphy MP, Dominiczak AF. 2009. Mitochondria-targeted antioxidant MitoQ10 improves endothelial function and attenuates cardiac hypertrophy. Hypertension. 54(2):322–328.
  • Green H, Ross G, Peacock J, Owen R, Yarnold J, Houlston R. 2002. Variation in the manganese superoxide dismutase gene (SOD2) is not a major cause of radiotherapy complications in breast cancer patients. Radiother Oncol. 63(2):213–216.
  • Grivennikova VG, Kareyeva AV, Vinogradov AD. 2018. Oxygen-dependence of mitochondrial ROS production as detected by Amplex Red assay. Redox Biol. 17:192–199.
  • Gromadzka G, Kruszyńska M, Wierzbicka D, Litwin T, Dzieżyc K, Wierzchowska-Ciok A, Chabik G, Członkowska A. 2015. Gene variants encoding proteins involved in antioxidant defense system and the clinical expression of Wilson disease. Liver Int. 35(1):215–222.
  • Gu Q, Zhao L, Ma YP, Liu JD. 2015. Contribution of mitochondrial function to exercise-induced attenuation of renal dysfunction in spontaneously hypertensive rats. Mol Cell Biochem. 406(1–2):217–225.
  • Guigni BA, Callahan DM, Tourville TW, Miller MS, Fiske B, Voigt T, Korwin-Mihavics B, Anathy V, Dittus K, Toth MJ. 2018. Skeletal muscle atrophy and dysfunction in breast cancer patients: role for chemotherapy-derived oxidant stress. Am J Physiol, Cell Physiol. 315(5):C744–C756.
  • Guo L, Li L, Wang W, Pan Z, Zhou Q, Wu Z. 2012. Mitochondrial reactive oxygen species mediates nicotine-induced hypoxia-inducible factor-1alpha expression in human non-small cell lung cancer cells. Biochim Biophys Acta. 1822(6):852–861.
  • Guo H, Seixas-Silva JA, Jr., Epperly MW, Gretton JE, Shin DM, Bar-Sagi D, Archer H, Greenberger JS. 2003. Prevention of radiation-induced oral cavity mucositis by plasmid/liposome delivery of the human manganese superoxide dismutase (SOD2) transgene. Radiat Res. 159(3):361–370.
  • Guo C, Sun L, Chen X, Zhang D. 2013. Oxidative stress, mitochondrial damage and neurodegenerative diseases. Neural Regen Res. 8(21):2003–2014.
  • Gurel A, Tomac N, Yilmaz HR, Tekedereli I, Akyol O, Armutcu F, Yuce H, Akin H, Ozcelik N, Elyas H. 2004. The Ala-9Val polymorphism in the mitochondrial targeting sequence (MTS) of the manganese superoxide dismutase gene is not associated with juvenile-onset asthma. Clin Biochem. 37(12):1117–1120.
  • Gurgul E, Lortz S, Tiedge M, Jorns A, Lenzen S. 2004. Mitochondrial catalase overexpression protects insulin-producing cells against toxicity of reactive oxygen species and proinflammatory cytokines. Diabetes. 53(9):2271–2280.
  • Guzy RD, Sharma B, Bell E, Chandel NS, Schumacker PT. 2008. 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. 28(2):718–731.
  • Haghighi SF, Salehi Z, Mohamad RS. 2014. Analysis of MnSOD and VEGF genes polymorphisms in Diabetic Retinopathy. Mol Biochem Diagnosis. 1:13–20.
  • Haghighi SF, Salehi Z, Sabouri MR, Abbasi N, Haghighi SF, Salehi Z, Sabouri MR, Abbasi N. 2015. Polymorphic variant of MnSOD A16V and risk of diabetic retinopathy. Mol Biol. 49(1):99–118.
  • Hamada T, Nakane T, Kimura T, Arisawa K, Yoneda K, Yamamoto T, Osaki T. 1999. Treatment of xerostomia with the bile secretion-stimulating drug anethole trithione: a clinical trial. Am J Med Sci. 318(3):146–151.
  • Hamanaka RB, Chandel NS. 2010. Mitochondrial reactive oxygen species regulate cellular signaling and dictate biological outcomes. Trends Biochem Sci. 35(9):505–513.
  • Hamed MO. 2017. Evaluation of a novel mitochondria-targeted anti-oxidant therapy for ischaemia-reperfusion injury in renal transplantation [Doctoral Thesis]. doi:10.17863/CAM.13853
  • Han YH, Buffolo M, Pires KM, Pei S, Scherer PE, Boudina S. 2016. Adipocyte-specific deletion of manganese superoxide dismutase protects from diet-induced obesity through increased mitochondrial uncoupling and biogenesis. Diabetes. 65(9):2639–2651.
  • Han SJ, Choi HS, Kim JI, Park JW, Park KM. 2018. IDH2 deficiency increases the liver susceptibility to ischemia-reperfusion injury via increased mitochondrial oxidative injury. Redox Biol. 14:142–153.
  • Han Z, Varadharaj S, Giedt RJ, Zweier JL, Szeto HH, Alevriadou BR. 2009. Mitochondria-derived reactive oxygen species mediate heme oxygenase-1 expression in sheared endothelial cells. J Pharmacol Exp Ther. 329(1):94–101.
  • Han Y, Xu X, Tang C, Gao P, Chen X, Xiong X, Yang M, Yang S, Zhu X, Yuan S, et al. 2018. Reactive oxygen species promote tubular injury in diabetic nephropathy: The role of the mitochondrial ros-txnip-nlrp3 biological axis. Redox Biol. 16:32–46.
  • Hao L, Sun Q, Zhong W, Zhang W, Sun X, Zhou Z. 2018. Mitochondria-targeted ubiquinone (MitoQ) enhances acetaldehyde clearance by reversing alcohol-induced posttranslational modification of aldehyde dehydrogenase 2: a molecular mechanism of protection against alcoholic liver disease. Redox Biol. 14:626–636.
  • Harris N, Costa V, MacLean M, Mollapour M, Moradas-Ferreira P, Piper PW. 2003. Mnsod overexpression extends the yeast chronological (G(0)) life span but acts independently of Sir2p histone deacetylase to shorten the replicative life span of dividing cells. Free Rad Biol Med. 34(12):1599–1606.
  • Hattori F, Murayama N, Noshita T, Oikawa S. 2003. Mitochondrial peroxiredoxin-3 protects hippocampal neurons from excitotoxic injury in vivo. J Neurochem. 86(4):860–868.
  • He Q, Harris N, Ren J, Han X. 2014. Mitochondria-targeted antioxidant prevents cardiac dysfunction induced by tafazzin gene knockdown in cardiac myocytes. Oxid Med Cell Longev. 2014:654198.
  • Heid ME, Keyel PA, Kamga C, Shiva S, Watkins SC, Salter RD. 2013. Mitochondrial reactive oxygen species induces NLRP3-dependent lysosomal damage and inflammasome activation. J Immunol. 191(10):5230–5238.
  • Hernandez-Saavedra D, McCord JM. 2009. Association of a new intronic polymorphism of the SOD2 gene (G1677T) with cancer. Cell Biochem Funct. 27(4):223–227.
  • Hill VM, O’Connor RM, Sissoko GB, Irobunda IS, Leong S, Canman JC, Stavropoulos N, Shirasu-Hiza M. 2018. A bidirectional relationship between sleep and oxidative stress in Drosophila. PLoS Biol. 16(7):e2005206.
  • Hiroi S, Harada H, Nishi H, Satoh M, Nagai R, Kimura A. 1999. Polymorphisms in the SOD2 and HLA-DRB1 genes are associated with nonfamilial idiopathic dilated cardiomyopathy in Japanese. Biochem Biophys Res Commun. 261(2):332–339.
  • Hiroi M, Nagahara Y, Miyauchi R, Misaki Y, Goda T, Kasezawa N, Sasaki S, Yamakawa-Kobayashi K. 2011. The combination of genetic variations in the PRDX3 gene and dietary fat intake contribute to obesity risk. Obesity (Silver Spring). 19(4):882–887.
  • Hitzeroth A, Niehaus DJ, Koen L, Botes WC, Deleuze JF, Warnich L. 2007. Association between the MnSOD Ala-9Val polymorphism and development of schizophrenia and abnormal involuntary movements in the Xhosa population. Prog Neuropsychopharmacol Biol Psychiatry. 31(3):664–672.
  • Ho YS, Magnenat JL, Bronson RT, Cao J, Gargano M, Sugawara M, Funk CD. 1997. Mice deficient in cellular glutathione peroxidase develop normally and show no increased sensitivity to hyperoxia. J Biol Chem. 272(26):16644–16651.
  • Hodge DR, Peng B, Pompeia C, Thomas S, Cho E, Clausen PA, Marquez VE, Farrar WL. 2005. Epigenetic silencing of manganese superoxide dismutase (SOD-2) in KAS 6/1 human multiple myeloma cells increases cell proliferation. Cancer Biol Ther. 4(5):585–592.
  • Hodge DR, Xiao W, Peng B, Cherry JC, Munroe DJ, Farrar WL. 2005. Enforced expression of superoxide dismutase 2/manganese superoxide dismutase disrupts autocrine interleukin-6 stimulation in human multiple myeloma cells and enhances dexamethasone-induced apoptosis. Cancer Res. 65(14):6255–6263.
  • Hoehn KL, Salmon AB, Hohnen-Behrens C, Turner N, Hoy AJ, Maghzal GJ, Stocker R, Van Remmen H, Kraegen EW, Cooney GJ, et al. 2009. Insulin resistance is a cellular antioxidant defense mechanism. Proc Natl Acad Sci USA. 106(42):17787–17792.
  • Hoffman DL, Salter JD, Brookes PS. 2007. Response of mitochondrial reactive oxygen species generation to steady-state oxygen tension: implications for hypoxic cell signaling. Am J Physiol Heart Circ Physiol. 292(1):H101–H108.
  • Holla LI, Kankova K, Vasku A. 2006. Functional polymorphism in the manganese superoxide dismutase (MnSOD) gene in patients with asthma. Clin Biochem. 39(3):299–302.
  • Holmstrom KM, Finkel T. 2014. Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat Rev Mol Cell Biol. 15(6):411–421.
  • Ho-Pun-Cheung A, Assenat E, Bascoul-Mollevi C, Bibeau F, Boissière-Michot F, Thezenas S, Cellier D, Azria D, Rouanet P, Senesse P, et al. 2011. A large-scale candidate gene approach identifies SNPs in SOD2 and IL13 as predictive markers of response to preoperative chemoradiation in rectal cancer. Pharmacogenomics J. 11(6):437–443.
  • Hori H, Ohmori O, Shinkai T, Kojima H, Okano C, Suzuki T, Nakamura J. 2000. Manganese superoxide dismutase gene polymorphism and schizophrenia: relation to tardive dyskinesia. Neuropsychopharmacology. 23(2):170–177.
  • Hoshino A, Okawa Y, Ariyoshi M, Kaimoto S, Uchihashi M, Fukai K, Iwai-Kanai E, Matoba S. 2014. Oxidative post-translational modifications develop LONP1 dysfunction in pressure overload heart failure. Circ Heart Fail. 7(3):500–509.
  • Hosoki A, Yonekura S, Zhao QL, Wei ZL, Takasaki I, Tabuchi Y, Wang LL, Hasuike S, Nomura T, Tachibana A, et al. 2012. Mitochondria-targeted superoxide dismutase (SOD2) regulates radiation resistance and radiation stress response in HeLa cells. J Radiat Res. 53(1):58–71.
  • Hou Y, Li S, Wu M, Wei J, Ren Y, Du C, Wu H, Han C, Duan H, Shi Y. 2016. Mitochondria-targeted peptide SS-31 attenuates renal injury via an antioxidant effect in diabetic nephropathy. Am J Physiol Renal Physiol. 310(6):F547–F559.
  • Houldsworth A, Hodgkinson A, Shaw S, Millward A, Demaine AG. 2015. Polymorphic differences in the SOD-2 gene may affect the pathogenesis of nephropathy in patients with diabetes and diabetic complications. Gene. 569(1):41–45.
  • Houldsworth A, Metzner M, Shaw S, Kaminski E, Demaine AG, Cramp ME. 2014. Polymorphic differences in SOD-2 may influence HCV viral clearance. J Med Virol. 86(6):941–947.
  • Houstis N, Rosen ED, Lander ES. 2006. Reactive oxygen species have a causal role in multiple forms of insulin resistance. Nature. 440(7086):944–948.
  • Hovnik T, Dolzan V, Bratina NU, Podkrajsek KT, Battelino T. 2009. Genetic polymorphisms in genes encoding antioxidant enzymes are associated with diabetic retinopathy in type 1 diabetes. Diabetes Care. 32(12):2258–2262.
  • Hoyt LR, Randall MJ, Ather JL, DePuccio DP, Landry CC, Qian X, Janssen-Heininger YM, van der Vliet A, Dixon AE, Amiel E, et al. 2017. Mitochondrial ROS induced by chronic ethanol exposure promote hyper-activation of the NLRP3 inflammasome. Redox Biol. 12:883–896.
  • Hu D, Cao P, Thiels E, Chu CT, Wu GY, Oury TD, Klann E. 2007. Hippocampal long-term potentiation, memory, and longevity in mice that overexpress mitochondrial superoxide dismutase. Neurobiol Learn Mem. 87(3):372–384.
  • Hu W, Dang XB, Wang G, Li S, Zhang YL. 2018. Peroxiredoxin-3 attenuates traumatic neuronal injury through preservation of mitochondrial function. Neurochem Int. 114:120–126.
  • Hu H, Li M. 2016. Mitochondria-targeted antioxidant mitotempo protects mitochondrial function against amyloid beta toxicity in primary cultured mouse neurons. Biochem Biophys Res Commun. 478(1):174–180.
  • Hu Q, Ren J, Li G, Wu J, Wu X, Wang G, Gu G, Ren H, Hong Z, Li J. 2018. The mitochondrially targeted antioxidant MitoQ protects the intestinal barrier by ameliorating mitochondrial DNA damage via the Nrf2/ARE signaling pathway. Cell Death Dis. 9(3):403.
  • Huang S, Dong R, Xu G, Liu J, Gao X, Yu S, Qie P, Gou G, Hu M, Wang Y, et al. 2020. Synthesis, characterization, and in vivo evaluation of desmethyl anethole trithione phosphate prodrug for ameliorating cerebral ischemia-reperfusion injury in rats. ACS Omega. 5(9):4595–4602.
  • Huang J, Li X, Li M, Li J, Xiao W, Ma W, Chen X, Liang X, Tang S, Luo Y. 2013. Mitochondria-targeted antioxidant peptide SS31 protects the retinas of diabetic rats. Curr Mol Med. 13(6):935–945.
  • Huang L, Lyu J, Liu QP, Chen C, Wang T. 2017. MnSOD Val16Ala polymorphism associated with retinopathy risk in diabetes: a PRISMA-compliant meta-analysis of case-control studies. Int J Ophthalmol. 10(4):639–645.
  • Hudson MB, Smuder AJ, Nelson WB, Wiggs MP, Shimkus KL, Fluckey JD, Szeto HH, Powers SK. 2015. Partial support ventilation and mitochondrial-targeted antioxidants protect against ventilator-induced decreases in diaphragm muscle protein synthesis. PLoS One. 10(9):e0137693.
  • Huh JY, Kim Y, Jeong J, Park J, Kim I, Huh KH, Kim YS, Woo HA, Rhee SG, Lee K-J, et al. 2012. Peroxiredoxin 3 is a key molecule regulating adipocyte oxidative stress, mitochondrial biogenesis, and adipokine expression. Antioxid Redox Signal. 16(3):229–243.
  • Hulsmans M, Van Dooren E, Holvoet P. 2012. Mitochondrial reactive oxygen species and risk of atherosclerosis. Curr Atheroscler Rep. 14(3):264–276.
  • Hung RJ, Boffetta P, Brennan P, Malaveille C, Gelatti U, Placidi D, Carta A, Hautefeuille A, Porru S. 2004. Genetic polymorphisms of MPO, COMT, MnSOD, NQO1, interactions with environmental exposures and bladder cancer risk. Carcinogenesis. 25(6):973–978.
  • Hurd TR, Collins Y, Abakumova I, Chouchani ET, Baranowski B, Fearnley IM, Prime TA, Murphy MP, James AM. 2012. Inactivation of pyruvate dehydrogenase kinase 2 by mitochondrial reactive oxygen species. J Biol Chem. 287(42):35153–35160.
  • Hussain A, Pooryasin A, Zhang M, Loschek LF, La Fortezza M, Friedrich AB, Blais CM, Ucpunar HK, Yepez VA, Lehmann M, et al. 2018. Inhibition of oxidative stress in cholinergic projection neurons fully rescues aging-associated olfactory circuit degeneration in Drosophila. Elife. 7:e32018.
  • Hwang I, Uddin MJ, Lee G, Jiang S, Pak ES, Ha H. 2019. Peroxiredoxin 3 deficiency accelerates chronic kidney injury in mice through interactions between macrophages and tubular epithelial cells. Free Radic Biol Med. 131:162–172.
  • Ibrahim AA, Karam HM, Shaaban EA, Safar MM, El-Yamany MF. 2019. MitoQ ameliorates testicular damage induced by gamma irradiation in rats: modulation of mitochondrial apoptosis and steroidogenesis. Life Sci. 232:116655.
  • Imlay JA. 2006. Iron-sulphur clusters and the problem with oxygen. Mol Microbiol. 59(4):1073–1082.
  • Iomdina EN, Khoroshilova-Maslova IP, Robustova OV, Averina OA, Kovaleva NA, Aliev G, Reddy VP, Zamyatnin AA, Jr., Skulachev MV, Senin II, et al. 2015. Mitochondria-targeted antioxidant SkQ1 reverses glaucomatous lesions in rabbits. Front Biosci (Landmark Ed). 20:892–901.
  • Iqbal A, Naz M, Hussain F, Saleem M, Hashmi BK. 2013. Association of SOD2 gene polymorphism with noise induced hearing loss in textile industry employees. African J Biochem Res. 7:122–127.
  • Isaev NK, Novikova SV, Stelmashook EV, Barskov IV, Silachev DN, Khaspekov LG, Skulachev VP, Zorov DB. 2012. Mitochondria-targeted plastoquinone antioxidant SkQR1 decreases trauma-induced neurological deficit in rat. Biochemistry Mosc. 77(9):996–999.
  • Isaev NK, Stelmashook EV, Genrikhs EE, Korshunova GA, Sumbatyan NV, Kapkaeva MR, Skulachev VP. 2016. Neuroprotective properties of mitochondria-targeted antioxidants of the SkQ-type. Rev Neurosci. 27(8):849–855.
  • Ishikawa K, Funayama T, Ohde H, Inagaki Y, Mashima Y. 2005. Genetic variants of TP53 and EPHX1 in Leber’s hereditary optic neuropathy and their relationship to age at onset. Jpn J Ophthalmol. 49(2):121–126.
  • Isikli A, Kubat-Uzum A, Satman I, Matur Z, Oge AE, Kucukali CI, Tuzun E, Erden S, Ozkok E. 2018. A SOD2 polymorphism is associated with abnormal quantitative sensory testing in type 2 diabetic patients. Noro Psikiyatr Ars. 55(3):276–279.
  • Itani HA, Dikalova AE, McMaster WG, Nazarewicz RR, Bikineyeva AT, Harrison DG, Dikalov SI. 2016. Mitochondrial cyclophilin D in vascular oxidative stress and hypertension. Hypertension. 67(6):1218–1227.
  • Jaffer OA, Carter AB, Sanders PN, Dibbern ME, Winters CJ, Murthy S, Ryan AJ, Rokita AG, Prasad AM, Zabner J, et al. 2015. Mitochondrial-targeted antioxidant therapy decreases transforming growth factor-beta-mediated collagen production in a murine asthma model. Am J Respir Cell Mol Biol. 52(1):106–115.
  • Jang S, Lewis TS, Powers C, Khuchua Z, Baines CP, Wipf P, Javadov S. 2017. Elucidating mitochondrial electron transport chain supercomplexes in the heart during ischemia-reperfusion. Antioxid Redox Signal. 27(1):57–69.
  • Jang YC, Perez VI, Song W, Lustgarten MS, Salmon AB, Mele J, Qi W, Liu Y, Liang H, Chaudhuri A, et al. 2009. Overexpression of Mn superoxide dismutase does not increase life span in mice. J Gerontol A Biol Sci Med Sci. 64A(11):1114–1125.
  • Jang S, Yu LR, Abdelmegeed MA, Gao Y, Banerjee A, Song BJ. 2015. Critical role of c-jun N-terminal protein kinase in promoting mitochondrial dysfunction and acute liver injury. Redox Biol. 6:552–564.
  • Jankauskas SS, Plotnikov EY, Morosanova MA, Pevzner IB, Zorova LD, Skulachev VP, Zorov DB. 2012. Mitochondria-targeted antioxidant SkQR1 ameliorates gentamycin-induced renal failure and hearing loss. Biochemistry Mosc. 77(6):666–670.
  • Jastroch M, Divakaruni AS, Mookerjee S, Treberg JR, Brand MD. 2010. Mitochondrial proton and electron leaks. Essays Biochem. 47:53–67.
  • Jauslin ML, Meier T, Smith RA, Murphy MP. 2003. Mitochondria-targeted antioxidants protect Friedreich Ataxia fibroblasts from endogenous oxidative stress more effectively than untargeted antioxidants. FASEB J. 17(13):1972–1974.
  • Jerotic D, Matic M, Suvakov S, Vucicevic K, Damjanovic T, Savic-Radojevic A, Pljesa-Ercegovac M, Coric V, Stefanovic A, Ivanisevic J, et al. 2019. Association of Nrf2, SOD2 and GPX1 polymorphisms with biomarkers of oxidative distress and survival in end-stage renal disease patients. Toxins (Basel). 11(7):431.
  • Jezek P, Hlavata L. 2005. Mitochondria in homeostasis of reactive oxygen species in cell, tissues, and organism. Int J Biochem Cell Biol. 37(12):2478–2503.
  • Ji G, Gu A, Wang Y, Huang C, Hu F, Zhou Y, Song L, Wang X. 2012. Genetic variants in antioxidant genes are associated with sperm DNA damage and risk of male infertility in a Chinese population. Free Rad Biol Med. 52(4):775–780.
  • Jia YL, Sun SJ, Chen JH, Jia Q, Huo TT, Chu LF, Bai JT, Yu YJ, Yan XX, Wang JH. 2016. SS31, a small molecule antioxidant peptide, attenuates beta-amyloid elevation, mitochondrial/synaptic deterioration and cognitive deficit in SAMP8 mice. Curr Alzheimer Res. 13(3):297–306.
  • Jiang MY, Han ZD, Li W, Yue F, Ye J, Li B, Cai Z, Lu JM, Dong W, Jiang X, et al. 2017. Mitochondrion-associated protein peroxiredoxin 3 promotes benign prostatic hyperplasia through autophagy suppression and pyroptosis activation. Oncotarget. 8(46):80295–80302.
  • Jiang B, Hebert VY, Li Y, Mathis JM, Alexander JS, Dugas TR. 2007. HIV antiretroviral drug combination induces endothelial mitochondrial dysfunction and reactive oxygen species production, but not apoptosis. Toxicol Appl Pharmacol. 224(1):60–71.
  • Jiang Y, Liu C, Lei B, Xu X, Lu B. 2018. Mitochondria-targeted antioxidant SkQ1 improves spermatogenesis in Immp2l mutant mice. Andrologia. 50(2):e12848.
  • Jiang J, Stoyanovsky DA, Belikova NA, Tyurina YY, Zhao Q, Tungekar MA, Kapralova V, Huang Z, Mintz AH, Greenberger JS, et al. 2009. A mitochondria-targeted triphenylphosphonium-conjugated nitroxide functions as a radioprotector/mitigator. Radiat Res. 172(6):706–717.
  • Jiang W, Tang L, Zeng J, Chen B. 2016. Adeno-associated virus mediated SOD gene therapy protects the retinal ganglion cells from chronic intraocular pressure elevation induced injury via attenuating oxidative stress and improving mitochondrial dysfunction in a rat model. Am J Transl Res. 8(2):799–810.
  • Jin H, Kanthasamy A, Ghosh A, Anantharam V, Kalyanaraman B, Kanthasamy AG. 2014. Mitochondria-targeted antioxidants for treatment of Parkinson’s disease: preclinical and clinical outcomes. Biochim Biophys Acta. 1842(8):1282–1294.
  • Johnson JM, Ferrara PJ, Verkerke ARP, Coleman CB, Wentzler EJ, Neufer PD, Kew KA, de Castro Bras LE, Funai K. 2018. Targeted overexpression of catalase to mitochondria does not prevent cardioskeletal myopathy in Barth syndrome. J Mol Cell Cardiol. 121:94–102.
  • Jones DA, Prior SL, Tang TS, Bain SC, Hurel SJ, Humphries SE, Stephens JW. 2010. Association between the rs4880 superoxide dismutase 2 (C>T) gene variant and coronary heart disease in diabetes mellitus. Diabetes Res Clin Pract. 90(2):196–201.
  • Jong CJ, Ito T, Schaffer SW. 2015. The ubiquitin-proteasome system and autophagy are defective in the taurine-deficient heart. Amino Acids. 47(12):2609–2622.
  • Jówko E, Gierczuk D, Cieśliński I, Kotowska J. 2017. SOD2 gene polymorphism and response of oxidative stress parameters in young wrestlers to a three-month training. Free Radic Res. 51(5):506–516.
  • Jówko E, Gromisz W, Sadowski J, Cieśliński I, Kotowska J. 2017. SOD2 gene polymorphism may modulate biochemical responses to a 12-week swimming training. Free Rad Biol Med. 113:571–579.
  • Justilien V, Pang JJ, Renganathan K, Zhan X, Crabb JW, Kim SR, Sparrow JR, Hauswirth WW, Lewin AS. 2007. SOD2 knockdown mouse model of early AMD. Invest Ophthalmol Vis Sci. 48(10):4407–4420.
  • Kakko S, Paivansalo M, Koistinen P, Kesaniemi YA, Kinnula VL, Savolainen MJ. 2003. The signal sequence polymorphism of the MnSOD gene is associated with the degree of carotid atherosclerosis. Atherosclerosis. 168(1):147–152.
  • Kalen AL, Sarsour EH, Venkataraman S, Goswami PC. 2006. Mn-superoxide dismutase overexpression enhances G2 accumulation and radioresistance in human oral squamous carcinoma cells. Antioxid Redox Signal. 8(7–8):1273–1281.
  • Kalyanaraman B, Hardy M, Podsiadly R, Cheng G, Zielonka J. 2017. Recent developments in detection of superoxide radical anion and hydrogen peroxide: Opportunities, challenges, and implications in redox signaling. Arch Biochem Biophys. 617:38–47.
  • Kalyanaraman B, Darley-Usmar V, Davies KJA, Dennery PA, Forman HJ, Grisham MB, Mann GE, Moore K, Roberts LJ, Ischiropoulos H. 2012. Measuring reactive oxygen and nitrogen species with fluorescent probes: challenges and limitations. Free Radic Biol Med. 52(1):1–6.
  • Kanai AJ, Zeidel ML, Lavelle JP, Greenberger JS, Birder LA, de Groat WC, Apodaca GL, Meyers SA, Ramage R, Epperly MW. 2002. Manganese superoxide dismutase gene therapy protects against irradiation-induced cystitis. Am J Physiol Renal Physiol. 283(6):F1304–F1312.
  • Kang PT, Chen CL, Ohanyan V, Luther DJ, Meszaros JG, Chilian WM, Chen YR. 2015. Overexpressing superoxide dismutase 2 induces a supernormal cardiac function by enhancing redox-dependent mitochondrial function and metabolic dilation. J Mol Cell Cardiol. 88:14–28.
  • Kang L, Liu S, Li J, Tian Y, Xue Y, Liu X. 2020. The mitochondria-targeted anti-oxidant MitoQ protects against intervertebral disc degeneration by ameliorating mitochondrial dysfunction and redox imbalance. Cell Prolif. 53(3):e12779.
  • Kangas-Kontio T, Vavuli S, Kakko SJ, Penna J, Savolainen ER, Savolainen MJ, Liinamaa MJ. 2009. Polymorphism of the manganese superoxide dismutase gene but not of vascular endothelial growth factor gene is a risk factor for diabetic retinopathy. Br J Ophthalmol. 93(10):1401–1406.
  • Kang L, Dai C, Lustig ME, Bonner JS, Mayes WH, Mokshagundam S, James FD, Thompson CS, Lin C-T, Perry CGR, et al. 2014. Heterozygous SOD2 deletion impairs glucose-stimulated insulin secretion, but not insulin action, in high-fat-fed mice. Diabetes. 63(11):3699–3710.
  • Kang D, Lee K-M, Park SK, Berndt SI, Peters U, Reding D, Chatterjee N, Welch R, Chanock S, Huang W-Y, et al. 2007. Functional variant of manganese superoxide dismutase (SOD2 V16A) polymorphism is associated with prostate cancer risk in the prostate, lung, colorectal, and ovarian cancer study. Cancer Epidemiol Biomarkers Prev. 16(8):1581–1586.
  • Kang L, Lustig ME, Bonner JS, Lee-Young RS, Mayes WH, James FD, Lin C-T, Perry CGR, Anderson EJ, Neufer PD, et al. 2012. Mitochondrial antioxidative capacity regulates muscle glucose uptake in the conscious mouse: effect of exercise and diet. J Appl Physiol. 113(8):1173–1183.
  • Kanwar M, Chan PS, Kern TS, Kowluru RA. 2007. Oxidative damage in the retinal mitochondria of diabetic mice: possible protection by superoxide dismutase. Invest Ophthalmol Vis Sci. 48(8):3805–3811.
  • Karaa A, Haas R, Goldstein A, Vockley J, Weaver WD, Cohen BH. 2018. Randomized dose-escalation trial of elamipretide in adults with primary mitochondrial myopathy. Neurology. 90(14):e1212–e1221.
  • Kasahara E, Lin LR, Ho YS, Reddy VN. 2005. SOD2 protects against oxidation-induced apoptosis in mouse retinal pigment epithelium: implications for age-related macular degeneration. Invest Ophthalmol Vis Sci. 46(9):3426–3434.
  • Kazemi E, Moradi MT, Yari K, Mousavi SA, Kahrizi D. 2015. Association between Manganese Superoxide Dismutase (MnSOD Val-9Ala) genotypes with the risk of generalized aggressive periodontitis disease. Cell Mol Biol (Noisy-le-Grand). 61(8):49–52.
  • Kegler A, Cardoso AS, Caprara ALF, Pascotini ET, Arend J, Gabbi P, Duarte M, da Cruz IBM, Furian AF, Oliveira MS, et al. 2019. Involvement of MnSOD Ala16Val polymorphism in epilepsy: A relationship with seizure type, inflammation, and metabolic syndrome. Gene. 711:143924.
  • Keller JN, Kindy MS, Holtsberg FW, St Clair DK, Yen HC, Germeyer A, Steiner SM, Bruce-Keller AJ, Hutchins JB, Mattson MP. 1998. Mitochondrial manganese superoxide dismutase prevents neural apoptosis and reduces ischemic brain injury: suppression of peroxynitrite production, lipid peroxidation, and mitochondrial dysfunction. J Neurosci. 18(2):687–697.
  • Kelly B, Tannahill GM, Murphy MP, O’Neill LA. 2015. Metformin inhibits the production of reactive oxygen species from NADH:ubiquinone oxidoreductase to limit induction of interleukin-1beta (IL-1beta) and boosts interleukin-10 (IL-10) in lipopolysaccharide (LPS)-activated macrophages. J Biol Chem. 290(33):20348–20359.
  • Kelso GF, Porteous CM, Coulter CV, Hughes G, Porteous WK, Ledgerwood EC, Smith RA, Murphy MP. 2001. Selective targeting of a redox-active ubiquinone to mitochondria within cells: antioxidant and antiapoptotic properties. J Biol Chem. 276(7):4588–4596.
  • Khabour OF, Abdelhalim ES, Abu-Wardeh A. 2009. Association between SOD2 T-9C and MTHFR C677T polymorphisms and longevity: a study in Jordanian population. BMC Geriatr. 9:57.
  • Khan AA, Paget JT, McLaughlin M, Kyula JN, Wilkinson MJ, Pencavel T, Mansfield D, Roulstone V, Seth R, Halle M, et al. 2018. Genetically modified lentiviruses that preserve microvascular function protect against late radiation damage in normal tissues. Sci Transl Med. 10(425):eaar2041.
  • Kidir V, Uz E, Yigit A, Altuntas A, Yigit B, Inal S, Uz E, Sezer MT, Yilmaz HR. 2016. Manganese superoxide dismutase, glutathione peroxidase and catalase gene polymorphisms and clinical outcomes in acute kidney injury. Ren Fail. 38(3):372–377.
  • Kim YR, Baek JI, Kim SH, Kim MA, Lee B, Ryu N, Kim KH, Choi DG, Kim HM, Murphy MP, et al. 2019. Therapeutic potential of the mitochondria-targeted antioxidant MitoQ in mitochondrial-ROS induced sensorineural hearing loss caused by Idh2 deficiency. Redox Biol. 20:544–555.
  • Kim SJ, Cheresh P, Jablonski RP, Morales-Nebreda L, Cheng Y, Hogan E, Yeldandi A, Chi M, Piseaux R, Ridge K, et al. 2016. Mitochondrial catalase overexpressed transgenic mice are protected against lung fibrosis in part via preventing alveolar epithelial cell mitochondrial DNA damage. Free Radic Biol Med. 101:482–490.
  • Kim JH, Choi TG, Park S, Yun HR, Nguyen NNY, Jo YH, Jang M, Kim J, Kim J, Kang I, et al. 2018. Mitochondrial ROS-derived PTEN oxidation activates PI3K pathway for mTOR-induced myogenic autophagy. Cell Death Differ. 25(11):1921–1937.
  • Kim A, Joseph S, Khan A, Epstein CJ, Sobel R, Huang TT. 2010. Enhanced expression of mitochondrial superoxide dismutase leads to prolonged in vivo cell cycle progression and up-regulation of mitochondrial thioredoxin. Free Radic Biol Med. 48(11):1501–1512.
  • Kim H, Kim SH, Cha H, Kim SR, Lee JH, Park JW. 2016. IDH2 deficiency promotes mitochondrial dysfunction and dopaminergic neurotoxicity: implications for Parkinson’s disease. Free Radic Res. 50(8):853–860.
  • Kim GW, Kondo T, Noshita N, Chan PH. 2002. Manganese superoxide dismutase deficiency exacerbates cerebral infarction after focal cerebral ischemia/reperfusion in mice: implications for the production and role of superoxide radicals. Stroke. 33(3):809–815.
  • Kim H, Lee YD, Kim HJ, Lee ZH, Kim HH. 2017. SOD2 and Sirt3 control osteoclastogenesis by regulating mitochondrial ROS. J Bone Miner Res. 32(2):397–406.
  • Kim H, Lee JH, Park JW. 2019. IDH2 deficiency exacerbates acetaminophen hepatotoxicity in mice via mitochondrial dysfunction-induced apoptosis. Biochim Biophys Acta Mol Basis Dis. 1865(9):2333–2341.
  • Kim SH, Park JW. 2019. IDH2 deficiency impairs cutaneous wound healing via ROS-dependent apoptosis. Biochim Biophys Acta Mol Basis Dis. 1865(11):165523.
  • Kim S, Song J, Ernst P, Latimer MN, Ha CM, Goh KY, Ma W, Rajasekaran NS, Zhang J, Liu XM, et al. 2020. MitoQ regulates redox-related non-coding RNAs to preserve mitochondrial network integrity in pressure overload heart failure. Am J Physiol Heart Circ Physiol. 318(3):H682–H695.
  • Kim M, Stepanova A, Niatsetskaya Z, Sosunov S, Arndt S, Murphy MP, Galkin A, Ten VS. 2018. Attenuation of oxidative damage by targeting mitochondrial complex I in neonatal hypoxic-ischemic brain injury. Free Radic Biol Med. 124:517–524.
  • Kim EH, Tolhurst AT, Szeto HH, Cho SH. 2015. Targeting CD36-mediated inflammation reduces acute brain injury in transient, but not permanent, ischemic stroke. CNS Neurosci Ther. 21(4):385–391.
  • Kim A, Zhong W, Oberley TD. 2004. Reversible modulation of cell cycle kinetics in NIH/3T3 mouse fibroblasts by inducible overexpression of mitochondrial manganese superoxide dismutase. Antioxid Redox Signal. 6(3):489–500.
  • Kimura K, Isashiki Y, Sonoda S, Kakiuchi-Matsumoto T, Ohba N. 2000. Genetic association of manganese superoxide dismutase with exudative age-related macular degeneration. Am J Ophthalmol. 130(6):769–773.
  • King A, Gottlieb E, Brooks DG, Murphy MP, Dunaief JL. 2004. Mitochondria-derived reactive oxygen species mediate blue light-induced death of retinal pigment epithelial cells. Photochem Photobiol. 79(5):470–475.
  • Kiningham KKS, Clair DK. 1997. Overexpression of manganese superoxide dismutase selectively modulates the activity of Jun-associated transcription factors in fibrosarcoma cells. Cancer Res. 57(23):5265–5271.
  • Kinnula VL, Lehtonen S, Koistinen P, Kakko S, Savolainen M, Kere J, Ollikainen V, Laitinen T. 2004. Two functional variants of the superoxide dismutase genes in Finnish families with asthma. Thorax. 59(2):116–119.
  • Kinoshita M, Sakamoto T, Kashio A, Shimizu T, Yamasoba T. 2013. Age-related hearing loss in Mn-SOD heterozygous knockout mice. Oxid Med Cell Longev. 2013:325702.
  • Kinugawa S, Wang Z, Kaminski PM, Wolin MS, Edwards JG, Kaley G, Hintze TH. 2005. Limited exercise capacity in heterozygous manganese superoxide dismutase gene-knockout mice: roles of superoxide anion and nitric oxide. Circulation. 111(12):1480–1486.
  • Kizhakekuttu TJ, Wang J, Dharmashankar K, Ying R, Gutterman DD, Vita JA, Widlansky ME. 2012. Adverse alterations in mitochondrial function contribute to type 2 diabetes mellitus-related endothelial dysfunction in humans. Arterioscler Thromb Vasc Biol. 32(10):2531–2539.
  • Klimova T, Chandel NS. 2008. Mitochondrial complex III regulates hypoxic activation of HIF. Cell Death Differ. 15(4):660–666.
  • Klivenyi P, St Clair D, Wermer M, Yen HC, Oberley T, Yang L, Beal MF. 1998. Manganese superoxide dismutase overexpression attenuates MPTP toxicity. Neurobiol Dis. 5(4):253–258.
  • Kloner RA, Hale SL, Dai W, Gorman RC, Shuto T, Koomalsingh KJ, Gorman JH, 3rd, Sloan RC, Frasier CR, Watson CA, et al. 2012. Reduction of ischemia/reperfusion injury with bendavia, a mitochondria-targeting cytoprotective Peptide. J Am Heart Assoc. 1(3):e001644.
  • Koç E, Olgar Y, Turan B. 2019. MitoTEMPO increases the gastrointestinal motility in aged rats. Cyprus J Med Sci. 4(1):24–27.
  • Kohler JJ, Cucoranu I, Fields E, Green E, He S, Hoying A, Russ R, Abuin A, Johnson D, Hosseini SH, et al. 2009. Transgenic mitochondrial superoxide dismutase and mitochondrially targeted catalase prevent antiretroviral-induced oxidative stress and cardiomyopathy. Lab Invest. 89(7):782–790.
  • Koike M, Nojiri H, Ozawa Y, Watanabe K, Muramatsu Y, Kaneko H, Morikawa D, Kobayashi K, Saita Y, Sasho T, et al. 2015. Mechanical overloading causes mitochondrial superoxide and SOD2 imbalance in chondrocytes resulting in cartilage degeneration. Sci Rep. 5:11722.
  • Kokoszka JE, Coskun P, Esposito LA, Wallace DC. 2001. Increased mitochondrial oxidative stress in the Sod2 (+/-) mouse results in the age-related decline of mitochondrial function culminating in increased apoptosis. Proc Natl Acad Sci USA. 98(5):2278–2283.
  • Kondo N, Bessho H, Honda S, Negi A. 2009. SOD2 gene polymorphisms in neovascular age-related macular degeneration and polypoidal choroidal vasculopathy. Mol Vis. 15:1819–1826.
  • Kong MJ, Bak SH, Han KH, Kim JI, Park JW, Park KM. 2019. Fragmentation of kidney epithelial cell primary cilia occurs by cisplatin and these cilia fragments are excreted into the urine. Redox Biol. 20:38–45.
  • Kong MJ, Han SJ, Kim JI, Park JW, Park KM. 2018. Mitochondrial NADP(+)-dependent isocitrate dehydrogenase deficiency increases cisplatin-induced oxidative damage in the kidney tubule cells. Cell Death Dis. 9(5):488.
  • Konig J, Ott C, Hugo M, Jung T, Bulteau AL, Grune T, Hohn A. 2017. Mitochondrial contribution to lipofuscin formation. Redox Biol. 11:673–681.
  • Kowalski M, Bielecka-Kowalska A, Oszajca K, Eusebio M, Jaworski P, Bartkowiak J, Szemraj J. 2010. Manganese superoxide dismutase (MnSOD) gene (Ala-9Val, Ile58Thr) polymorphism in patients with age-related macular degeneration (AMD). Med Sci Monit. 16(4):CR190–CR196.
  • Kowaltowski AJ, de Souza-Pinto NC, Castilho RF, Vercesi AE. 2009. Mitochondria and reactive oxygen species. Free Radic Biol Med. 47(4):333–343.
  • Kowluru RA, Kowluru V, Xiong Y, Ho YS. 2006. Overexpression of mitochondrial superoxide dismutase in mice protects the retina from diabetes-induced oxidative stress. Free Radic Biol Med. 41(8):1191–1196.
  • Kroller-Schon S, Steven S, Kossmann S, Scholz A, Daub S, Oelze M, Xia N, Hausding M, Mikhed Y, Zinssius E, et al. 2014. Molecular mechanisms of the crosstalk between mitochondria and NADPH oxidase through reactive oxygen species-studies in white blood cells and in animal models. Antioxid Redox Signal. 20(2):247–266.
  • Ku HJ, Ahn Y, Lee JH, Park KM, Park JW. 2015. IDH2 deficiency promotes mitochondrial dysfunction and cardiac hypertrophy in mice. Free Radic Biol Med. 80:84–92.
  • Ku HJ, Park JW. 2017. Downregulation of IDH2 exacerbates H2O2-mediated cell death and hypertrophy. Redox Rep. 22(1):35–41.
  • Ku HJ, Park JH, Kim SH, Park JW. 2018. Isocitrate dehydrogenase 2 deficiency exacerbates dermis damage by ultraviolet-B via DeltaNp63 downregulation. Biochim Biophys Acta Mol Basis Dis. 1864(4 Pt A):1138–1147.
  • Kucukgergin C, Sanli O, Amasyali AS, Tefik T, Seckin S. 2012. Genetic variants of MnSOD and GPX1 and susceptibility to bladder cancer in a Turkish population. Med Oncol. 29(3):1928–1934.
  • Kucukgergin C, Sanli O, Tefik T, Aydin M, Ozcan F, Seckin S. 2012. Increased risk of advanced prostate cancer associated with MnSOD Ala-9-Val gene polymorphism. Mol Biol Rep. 39(1):193–198.
  • Kumar A, Davuluri G, Welch N, Kim A, Gangadhariah M, Allawy A, Priyadarshini A, McMullen MR, Sandlers Y, Willard B, et al. 2019. Oxidative stress mediates ethanol-induced skeletal muscle mitochondrial dysfunction and dysregulated protein synthesis and autophagy. Free Radic Biol Med. 145:284–299.
  • Kurundkar D, Kurundkar AR, Bone NB, Becker EJ, Jr., Liu W, Chacko B, Darley-Usmar V, Zmijewski JW, Thannickal VJ. 2019. SIRT3 diminishes inflammation and mitigates endotoxin-induced acute lung injury. JCI Insight. 4(1):e120722.
  • Kussmaul L, Hirst J. 2006. The mechanism of superoxide production by NADH:ubiquinone oxidoreductase (complex I) from bovine heart mitochondria. Proc Natl Acad Sci USA. 103(20):7607–7612.
  • Kuwahara H, Horie T, Ishikawa S, Tsuda C, Kawakami S, Noda Y, Kaneko T, Tahara S, Tachibana T, Okabe M, et al. 2010. Oxidative stress in skeletal muscle causes severe disturbance of exercise activity without muscle atrophy. Free Radic Biol Med. 48(9):1252–1262.
  • Kwak HB, Lee Y, Kim JH, Van Remmen H, Richardson AG, Lawler JM. 2015. MnSOD overexpression reduces fibrosis and pro-apoptotic signaling in the aging mouse heart. J Gerontol A Biol Sci Med Sci. 70(5):533–544.
  • Laddha NC, Dwivedi M, Gani AR, Shajil EM, Begum R. 2013. Involvement of superoxide dismutase isoenzymes and their genetic variants in progression of and higher susceptibility to vitiligo. Free Radic Biol Med. 65:1110–1125.
  • Lai Y-C, Lin S-K, Huang M-C, Liu H-C, Chiou Y-L, Chen C-H, Chen WJ. 2019. Relations of genetic variants in superoxide dismutase 2 and dystrobrevin-binding protein 1 to methamphetamine psychosis among methamphetamine dependents in Taiwan. Taiwan J Psychiatry. 33(2):83–91.
  • Laitano O, Ahn B, Patel N, Coblentz PD, Smuder AJ, Yoo JK, Christou DD, Adhihetty PJ, Ferreira LF. 2016. Pharmacological targeting of mitochondrial reactive oxygen species counteracts diaphragm weakness in chronic heart failure. J Appl Physiol. 120(7):733–742.
  • Lam S, MacAulay C, Le Riche JC, Dyachkova Y, Coldman A, Guillaud M, Hawk E, Christen MO, Gazdar AF. 2002. A randomized phase IIb trial of anethole dithiolethione in smokers with bronchial dysplasia. J Natl Cancer Inst. 94(13):1001–1009.
  • Lambert AJ, Brand MD. 2004. Superoxide production by NADH:ubiquinone oxidoreductase (complex I) depends on the pH gradient across the mitochondrial inner membrane. Biochem J. 382(2):511–517.
  • Lambert AJ, Brand MD. 2009. Reactive oxygen species production by mitochondria. Methods Mol Biol. 554:165–181.
  • Lambert AJ, Buckingham JA, Boysen HM, Brand MD. 2010. Low complex I content explains the low hydrogen peroxide production rate of heart mitochondria from the long-lived pigeon, Columba Livia. Aging Cell. 9(1):78–91.
  • Lan C, Tam PO, Chiang SW, Chan CK, Luk FO, Lee GK, Ngai JW, Law JS, Lam DS, Pang CP, et al. 2009. Manganese superoxide dismutase and chemokine genes polymorphisms in Chinese patients with anterior uveitis. Invest Ophthalmol Vis Sci. 50(12):5596–5600.
  • Landi S, Gemignani F, Neri M, Barale R, Bonassi S, Bottari F, Canessa PA, Canzian F, Ceppi M, Filiberti R, et al. 2007. Polymorphisms of glutathione-S-transferase M1 and manganese superoxide dismutase are associated with the risk of malignant pleural mesothelioma. Int J Cancer. 120(12):2739–2743.
  • Lark DS, Kang L, Lustig ME, Bonner JS, James FD, Neufer PD, Wasserman DH. 2015. Enhanced mitochondrial superoxide scavenging does not improve muscle insulin action in the high fat-fed mouse. PLoS One. 10(5):e0126732.
  • Lassen O, Tabares S, Ojeda S, Dotto G, Bertolotto P, Sembaj A. 2018. Genetic polymorphisms of manganese-dependent superoxide dismutase in Chagas disease. Infect Dis Clin Pract. 26:159–164.
  • Lawana V, Singh N, Sarkar S, Charli A, Jin H, Anantharam V, Kanthasamy AG, Kanthasamy A. 2017. Involvement of c-Abl kinase in microglial activation of NLRP3 inflammasome and impairment in autolysosomal system. J Neuroimmune Pharmacol. 12(4):624–660.
  • Lebovitz RM, Zhang H, Vogel H, Cartwright J, Jr., Dionne L, Lu N, Huang S, Matzuk MM. 1996. Neurodegeneration, myocardial injury, and perinatal death in mitochondrial superoxide dismutase-deficient mice. Proc Natl Acad Sci Usa. 93(18):9782–9787.
  • Lee CY, Chang CH, Teng NC, Chang HM, Huang WT, Huang YK. 2019. Associations between the phenotype and genotype of MnSOD and catalase in periodontal disease. BMC Oral Health. 19(1):201.
  • Lee HY, Choi CS, Birkenfeld AL, Alves TC, Jornayvaz FR, Jurczak MJ, Zhang D, Woo DK, Shadel GS, Ladiges W, et al. 2010. Targeted expression of catalase to mitochondria prevents age-associated reductions in mitochondrial function and insulin resistance. Cell Metab. 12(6):668–674.
  • Lee HY, Kaneki M, Andreas J, Tompkins RG, Martyn JA. 2011. Novel mitochondria-targeted antioxidant peptide ameliorates burn-induced apoptosis and endoplasmic reticulum stress in the skeletal muscle of mice. Shock. 36(6):580–585.
  • Lee SR, Kim JR, Kwon KS, Yoon HW, Levine RL, Ginsburg A, Rhee SG. 1999. Molecular cloning and characterization of a mitochondrial selenocysteine-containing thioredoxin reductase from rat liver. J Biol Chem. 274(8):4722–4734.
  • Lee HY, Lee JS, Alves T, Ladiges W, Rabinovitch PS, Jurczak MJ, Choi CS, Shulman GI, Samuel VT. 2017. Mitochondrial-targeted catalase protects against high-fat diet-induced muscle insulin resistance by decreasing intramuscular lipid accumulation. Diabetes. 66(8):2072–2081.
  • Lee KP, Shin YJ, Cho SC, Lee SM, Bahn YJ, Kim JY, Kwon ES, Jeong DY, Park SC, Rhee SG, et al. 2014. Peroxiredoxin 3 has a crucial role in the contractile function of skeletal muscle by regulating mitochondrial homeostasis. Free Radic Biol Med. 77:298–306.
  • Lee S, Tak E, Lee J, Rashid MA, Murphy MP, Ha J, Kim SS. 2011. Mitochondrial H2O2 generated from electron transport chain complex I stimulates muscle differentiation. Cell Res. 21(5):817–834.
  • Lee SY, Usui S, Zafar AB, Oveson BC, Jo YJ, Lu L, Masoudi S, Campochiaro PA. 2011. N-Acetylcysteine promotes long-term survival of cones in a model of retinitis pigmentosa. J Cell Physiol. 226(7):1843–1849.
  • Lee S, Van Remmen H, Csete M. 2009. Sod2 overexpression preserves myoblast mitochondrial mass and function, but not muscle mass with aging. Aging Cell. 8(3):296–310.
  • Lee S, Wi SM, Min Y, Lee KY. 2016. Peroxiredoxin-3 Is involved in bactericidal activity through the regulation of mitochondrial reactive oxygen species. Immune Netw. 16(6):373–380.
  • Lehmann TG, Wheeler MD, Froh M, Schwabe RF, Bunzendahl H, Samulski RJ, Lemasters JJ, Brenner DA, Thurman RG. 2003. Effects of three superoxide dismutase genes delivered with an adenovirus on graft function after transplantation of fatty livers in the rat. Transplantation. 76(1):28–37.
  • Lenaz G. 2001. The mitochondrial production of reactive oxygen species: mechanisms and implications in human pathology. IUBMB Life. 52(3-5):159–164.
  • Lewen A, Fujimura M, Sugawara T, Matz P, Copin JC, Chan PH. 2001. Oxidative stress-dependent release of mitochondrial cytochrome c after traumatic brain injury. J Cereb Blood Flow Metab. 21(8):914–920.
  • Li F, Calingasan NY, Yu F, Mauck WM, Toidze M, Almeida CG, Takahashi RH, Carlson GA, Flint Beal M, Lin MT, et al. 2004. Increased plaque burden in brains of APP mutant MnSOD heterozygous knockout mice. J Neurochem. 89(5):1308–1312.
  • Li G, Chan YL, Sukjamnong S, Anwer AG, Vindin H, Padula M, Zakarya R, George J, Oliver BG, Saad S, et al. 2019. A mitochondrial specific antioxidant reverses metabolic dysfunction and fatty liver induced by maternal cigarette smoke in mice. Nutrients. 11(7):1669.
  • Li L, Chen Y, Gibson SB. 2013. Starvation-induced autophagy is regulated by mitochondrial reactive oxygen species leading to AMPK activation. Cell Signal. 25(1):50–65.
  • Li Y, Huang TT, Carlson EJ, Melov S, Ursell PC, Olson JL, Noble LJ, Yoshimura MP, Berger C, Chan PH, et al. 1995. Dilated cardiomyopathy and neonatal lethality in mutant mice lacking manganese superoxide dismutase. Nat Genet. 11(4):376–381.
  • Li L, Kaifu T, Obinata M, Takai T. 2009. Peroxiredoxin III-deficiency sensitizes macrophages to oxidative stress. J Biochem. 145(4):425–427.
  • Li D, Lai Y, Yue Y, Rabinovitch PS, Hakim C, Duan D. 2009. Ectopic catalase expression in mitochondria by adeno-associated virus enhances exercise performance in mice. PLoS One. 4(8):e6673.
  • Li XD, Li YM, Liu B, Guo X, Li YS. 2009. Association between SNPs in SOD2 and noise-induced hearing loss. Zhonghua Lao Dong Wei Sheng Zhi Ye Bing Za Zhi. 27(4):211–214.
  • Li JJ, Oberley LW. 1997. Overexpression of manganese-containing superoxide dismutase confers resistance to the cytotoxicity of tumor necrosis factor alpha and/or hyperthermia. Cancer Res. 57(10):1991–1998.
  • Li N, Oberley TD, Oberley LW, Zhong W. 1998a. Inhibition of cell growth in NIH/3T3 fibroblasts by overexpression of manganese superoxide dismutase: mechanistic studies. J Cell Physiol. 175(3):359–369.
  • Li N, Oberley TD, Oberley LW, Zhong W. 1998b. Overexpression of manganese superoxide dismutase in DU145 human prostate carcinoma cells has multiple effects on cell phenotype. Prostate. 35(3):221–233.
  • Li JJ, Oberley LW, St Clair DK, Ridnour LA, Oberley TD. 1995. Phenotypic changes induced in human breast cancer cells by overexpression of manganese-containing superoxide dismutase. Oncogene. 10(10):1989–2000.
  • Li L, Obinata M, Hori K. 2010. Role of peroxiredoxin III in the pathogenesis of pre-eclampsia as evidenced in mice. Oxid Med Cell Longev. 3(1):71–73.
  • Li R, Ren T, Zeng J. 2019. Mitochondrial Coenzyme Q protects sepsis-induced acute lung injury by activating PI3K/Akt/GSK-3beta/mTOR pathway in rats. Biomed Res Int. 2019:5240898.
  • Li L, Shoji W, Oshima H, Obinata M, Fukumoto M, Kanno N. 2008. Crucial role of peroxiredoxin III in placental antioxidant defense of mice. FEBS Lett. 582(16):2431–2434.
  • Li L, Shoji W, Takano H, Nishimura N, Aoki Y, Takahashi R, Goto S, Kaifu T, Takai T, Obinata M. 2007. Increased susceptibility of MER5 (peroxiredoxin III) knockout mice to LPS-induced oxidative stress. Biochem Biophys Res Commun. 355(3):715–721.
  • Li X, Song P, Timofeeva M, Meng X, Rudan I, Little J, Satsangi J, Campbell H, Theodoratou E. 2016. Systematic meta-analyses and field synopsis of genetic and epigenetic studies in paediatric inflammatory bowel disease. Sci Rep. 6:34076.
  • Li C, Sun H, Xu G, McCarter KD, Li J, Mayhan WG. 2018. 2018. Mito-Tempo prevents nicotine-induced exacerbation of ischemic brain damage. J Appl Physiol. 125(1):49–57.
  • Li X, Wang X, Zheng M, Luan QX. 2016. Mitochondrial reactive oxygen species mediate the lipopolysaccharide-induced pro-inflammatory response in human gingival fibroblasts. Exp Cell Res. 347(1):212–221.
  • Li G, Wu J, Li R, Yuan D, Fan Y, Yang J, Ji M, Zhu S. 2016. Protective effects of antioxidant peptide SS-31 against multiple organ dysfunctions during endotoxemia. Inflammation. 39(1):54–64.
  • Li F, Xu M, Wang M, Wang L, Wang H, Zhang H, Chen Y, Gong J, Zhang JJ, Adcock IM, et al. 2018. Roles of mitochondrial ROS and NLRP3 inflammasome in multiple ozone-induced lung inflammation and emphysema. Respir Res. 19(1):230.
  • Li FC, Yen JC, Chan SH, Chang AY. 2012. Bioenergetics failure and oxidative stress in brain stem mediates cardiovascular collapse associated with fatal methamphetamine intoxication. PLoS One. 7(1):e30589.
  • Lian K, Wang Q, Zhao S, Yang M, Chen G, Chen Y, Li C, Gao H, Li C. 2019. Pretreatment of diabetic adipose-derived stem cells with mitoTEMPO reverses their defective proangiogenic function in diabetic mice with critical limb ischemia. Cell Transplant. 28(12):1652–1663.
  • Liang LP, Ho YS, Patel M. 2000. Mitochondrial superoxide production in kainate-induced hippocampal damage. Neuroscience. 101(3):563–570.
  • Liang LP, Patel M. 2004. Mitochondrial oxidative stress and increased seizure susceptibility in Sod2(-/+) mice. Free Radic Biol Med. 36(5):542–554.
  • Liang HL, Sedlic F, Bosnjak Z, Nilakantan V. 2010. SOD1 and MitoTEMPO partially prevent mitochondrial permeability transition pore opening, necrosis, and mitochondrial apoptosis after ATP depletion recovery. Free Radic Biol Med. 49(10):1550–1560.
  • Liang LP, Waldbaum S, Rowley S, Huang TT, Day BJ, Patel M. 2012. Mitochondrial oxidative stress and epilepsy in SOD2 deficient mice: attenuation by a lipophilic metalloporphyrin. Neurobiol Dis. 45(3):1068–1076.
  • Liao AC, Craver BM, Tseng BP, Tran KK, Parihar VK, Acharya MM, Limoli CL. 2013. Mitochondrial-targeted human catalase affords neuroprotection from proton irradiation. Radiat Res. 180(1):1–6.
  • Liao Y, Zhang H, He D, Wang Y, Cai B, Chen J, Ma J, Liu Z, Wu Y. 2019. Retinal pigment epithelium cell death Is associated with NLRP3 inflammasome activation by all-trans retinal. Invest Ophthalmol Vis Sci. 60(8):3034–3045.
  • Lim S, Rashid MA, Jang M, Kim Y, Won H, Lee J, Woo JT, Kim YS, Murphy MP, Ali L, et al. 2011. Mitochondria-targeted antioxidants protect pancreatic beta-cells against oxidative stress and improve insulin secretion in glucotoxicity and glucolipotoxicity. Cell Physiol Biochem. 28(5):873–886.
  • Lin P, Hsueh YM, Ko JL, Liang YF, Tsai KJ, Chen CY. 2003. Analysis of NQO1, GSTP1, and MnSOD genetic polymorphisms on lung cancer risk in Taiwan. Lung Cancer. 40(2):123–129.
  • Lin Q, Li S, Jiang N, Shao X, Zhang M, Jin H, Zhang Z, Shen J, Zhou Y, Zhou W, et al. 2019. PINK1-parkin pathway of mitophagy protects against contrast-induced acute kidney injury via decreasing mitochondrial ROS and NLRP3 inflammasome activation. Redox Biol. 26:101254.
  • Liu C, Fang J, Liu W. 2019. Superoxide dismutase coding of gene polymorphisms associated with susceptibility to Parkinson’s disease. J Integr Neurosci. 18(3):299–303.
  • Liu Z, He Q, Ding X, Zhao T, Zhao L, Wang A. 2015. SOD2 is a C-myc target gene that promotes the migration and invasion of tongue squamous cell carcinoma involving cancer stem-like cells. Int J Biochem Cell Biol. 60:139–146.
  • Liu Z, Hu Y, Liang H, Sun Z, Feng S, Deng H. 2016. Silencing PRDX3 inhibits growth and promotes invasion and extracellular matrix degradation in hepatocellular carcinoma cells. J Proteome Res. 15(5):1506–1514.
  • Liu YM, Li XD, Guo X, Liu B, Lin AH, Ding YL, Rao SQ. 2010. SOD2 V16A SNP in the mitochondrial targeting sequence is associated with noise induced hearing loss in Chinese workers. Dis Markers. 28(3):137–147.
  • Liu Q, Liu H, Bai H, Huang W, Zhang R, Tan J, Guan L, Fan P. 2019. Association of SOD2 A16V and PON2 S311C polymorphisms with polycystic ovary syndrome in Chinese women. J Endocrinol Invest. 42(8):909–921.
  • Liu X, Murphy MP, Xing W, Wu H, Zhang R, Sun H. 2018. Mitochondria-targeted antioxidant MitoQ reduced renal damage caused by ischemia-reperfusion injury in rodent kidneys: Longitudinal observations of T2 -weighted imaging and dynamic contrast-enhanced MRI. Magn Reson Med. 79(3):1559–1567.
  • Liu R, Oberley TD, Oberley LW. 1997. Transfection and expression of MnSOD cDNA decreases tumor malignancy of human oral squamous carcinoma SCC-25 cells. Hum Gene Ther. 8(5):585–595.
  • Liu Y, Qi W, Richardson A, Van Remmen H, Ikeno Y, Salmon AB. 2013. Oxidative damage associated with obesity is prevented by overexpression of CuZn- or Mn-superoxide dismutase. Biochem Biophys Res Commun. 438(1):78–83.
  • Liu Y, Shim E, Nguyen P, Gibbons AT, Mitchell JB, Poirier MC. 2014. Tempol protects cardiomyocytes from nucleoside reverse transcriptase inhibitor-induced mitochondrial toxicity. Toxicol Sci. 139(1):133–141.
  • Liu Y, Tang L, Chen B. 2012. Effects of antioxidant gene therapy on retinal neurons and oxidative stress in a model of retinal ischemia/reperfusion. Free Radic Biol Med. 52(5):909–915.
  • Liu L, Wang M-j Y. T-h, Cheng Z, Li M, Guo Q-w. 2016. Mitochondria-targeted antioxidant Mitoquinone protects post-thaw human sperm against oxidative stress injury. Zhonghua nan ke xue. 22(3):205–211.
  • Liu Y, Wang Y, Ding W, Wang Y. 2018. Mito-TEMPO alleviates renal fibrosis by reducing inflammation, mitochondrial dysfunction, and endoplasmic reticulum stress. Oxid Med Cell Longev. 2018:5828120.
  • Liu G, Zhou W, Park S, Wang LI, Miller DP, Wain JC, Lynch TJ, Su L, Christiani DC. 2004. The SOD2 Val/Val genotype enhances the risk of nonsmall cell lung carcinoma by p53 and XRCC1 polymorphisms. Cancer. 101(12):2802–2808.
  • Liu G, Zhou W, Wang LI, Park S, Miller DP, Xu LL, Wain JC, Lynch TJ, Su L, Christiani DC. 2004. MPO and SOD2 polymorphisms, gender, and the risk of non-small cell lung carcinoma. Cancer Lett. 214(1):69–79.
  • Loor G, Kondapalli J, Iwase H, Chandel NS, Waypa GB, Guzy RD, Vanden Hoek TL, Schumacker PT. 2011. Mitochondrial oxidant stress triggers cell death in simulated ischemia-reperfusion. Biochim Biophys Acta. 1813(7):1382–1394.
  • Lopez-Acosta O, de Los Angeles Fortis-Barrera M, Barrios-Maya MA, Ramirez AR, Aguilar FJA, El-Hafidi M. 2018. Reactive oxygen species from NADPH oxidase and mitochondria participate in the proliferation of aortic smooth muscle cells from a model of metabolic syndrome. Oxid Med Cell Longev. 2018:1–10.
  • Lortz S, Gurgul-Convey E, Naujok O, Lenzen S. 2013. Overexpression of the antioxidant enzyme catalase does not interfere with the glucose responsiveness of insulin-secreting INS-1E cells and rat islets. Diabetologia. 56(4):774–782.
  • Lotteau S, Ivarsson N, Yang Z, Restagno D, Colyer J, Hopkins P, Weightman A, Himori K, Yamada T, Bruton J, et al. 2019. A mechanism for statin-induced susceptibility to myopathy. JACC Basic Transl Sci. 4(4):509–523.
  • Lowes DA, Thottakam BM, Webster NR, Murphy MP, Galley HF. 2008. The mitochondria-targeted antioxidant MitoQ protects against organ damage in a lipopolysaccharide-peptidoglycan model of sepsis. Free Radic Biol Med. 45(11):1559–1565.
  • Lowes DA, Wallace C, Murphy MP, Webster NR, Galley HF. 2009. The mitochondria targeted antioxidant MitoQ protects against fluoroquinolone-induced oxidative stress and mitochondrial membrane damage in human Achilles tendon cells. Free Radic Res. 43(4):323–328.
  • Lowes DA, Webster NR, Murphy MP, Galley HF. 2013. Antioxidants that protect mitochondria reduce interleukin-6 and oxidative stress, improve mitochondrial function, and reduce biochemical markers of organ dysfunction in a rat model of acute sepsis. Br J Anaesth. 110(3):472–480.
  • Lu X, Zhang Y, Bai H, Liu J, Li J, Wu B. 2018. Mitochondria-targeted antioxidant MitoTEMPO improves the post-thaw sperm quality. Cryobiology. 80:26–29.
  • Lubet RA, Steele VE, Eto I, Juliana MM, Kelloff GJ, Grubbs CJ. 1997. Chemopreventive efficacy of anethole trithione, N-acetyl-L-cysteine, miconazole and phenethylisothiocyanate in the DMBA-induced rat mammary cancer model. Int J Cancer. 72(1):95–101.
  • Lucena MI, Garcia-Martin E, Andrade RJ, Martinez C, Stephens C, Ruiz JD, Ulzurrun E, Fernandez MC, Romero-Gomez M, Castiella A, et al. 2010. Mitochondrial superoxide dismutase and glutathione peroxidase in idiosyncratic drug-induced liver injury. Hepatology. 52(1):303–312.
  • Lui G, Bouazza N, Denoyelle F, Moine M, Brugieres L, Chastagner P, Corradini N, Entz-Werle N, Verite C, Landmanparker J, et al. 2018. Association between genetic polymorphisms and platinum-induced ototoxicity in children. Oncotarget. 9(56):30883–30893.
  • Luo M, Guan X, Luczak ED, Lang D, Kutschke W, Gao Z, Yang J, Glynn P, Sossalla S, Swaminathan PD, et al. 2013. Diabetes increases mortality after myocardial infarction by oxidizing CaMKII. J Clin Invest. 123(3):1262–1274.
  • Luo ZC, Julien P, Wei SQ, Audibert F, Fraser WD, the Maternal and Infant Research on Oxidative Stress (MIROS) study group. 2018. Association of pre-eclampsia with SOD2 Ala16Val polymorphism among mother-father-infant triads. Int J Gynecol Obstet. 142(2):221–227.
  • Luo M, Shang L, Brooks MD, Jiagge E, Zhu Y, Buschhaus JM, Conley S, Fath MA, Davis A, Gheordunescu E, et al. 2018. Targeting breast cancer stem cell state equilibrium through modulation of redox signaling. Cell Metab. 28(1):69–86.e66.
  • Luo JD, Wang YY, Fu WL, Wu J, Chen AF. 2004. Gene therapy of endothelial nitric oxide synthase and manganese superoxide dismutase restores delayed wound healing in type 1 diabetic mice. Circulation. 110(16):2484–2493.
  • Luptak I, Qin F, Sverdlov AL, Pimentel DR, Panagia M, Croteau D, Siwik DA, Bachschmid MM, He H, Balschi JA, et al. 2019. Energetic dysfunction is mediated by mitochondrial reactive oxygen species and precedes structural remodeling in metabolic heart disease. Antioxid Redox Signal. 31(7):539–549.
  • Lustgarten MS, Jang YC, Liu Y, Muller FL, Qi W, Steinhelper M, Brooks SV, Larkin L, Shimizu T, Shirasawa T, et al. 2009. Conditional knockout of Mn-SOD targeted to type IIB skeletal muscle fibers increases oxidative stress and is sufficient to alter aerobic exercise capacity. Am J Physiol, Cell Physiol. 297(6):C1520–C1532.
  • Lustgarten MS, Jang YC, Liu Y, Qi W, Qin Y, Dahia PL, Shi Y, Bhattacharya A, Muller FL, Shimizu T, et al. 2011. MnSOD deficiency results in elevated oxidative stress and decreased mitochondrial function but does not lead to muscle atrophy during aging. Aging Cell. 10(3):493–505.
  • Ma T, Hoeffer CA, Wong H, Massaad CA, Zhou P, Iadecola C, Murphy MP, Pautler RG, Klann E. 2011. Amyloid beta-induced impairments in hippocampal synaptic plasticity are rescued by decreasing mitochondrial superoxide. J Neurosci. 31(15):5589–5595.
  • Macias CA, Chiao JW, Xiao J, Arora DS, Tyurina YY, Delude RL, Wipf P, Kagan VE, Fink MP. 2007. Treatment with a novel hemigramicidin-TEMPO conjugate prolongs survival in a rat model of lethal hemorrhagic shock. Ann Surg. 245(2):305–314.
  • Maier CM, Hsieh L, Crandall T, Narasimhan P, Chan PH. 2006. Evaluating therapeutic targets for reperfusion-related brain hemorrhage. Ann Neurol. 59(6):929–938.
  • Mailloux RJ. 2018. Mitochondrial antioxidants and the maintenance of cellular hydrogen peroxide levels. Oxid Med Cell Longev. 2018:7857251.
  • Maiti AK, Saha NC, More SS, Panigrahi AK, Paul G. 2017. Neuroprotective efficacy of mitochondrial antioxidant MitoQ in suppressing peroxynitrite-mediated mitochondrial dysfunction inflicted by lead toxicity in the rat rrain. Neurotox Res. 31(3):358–372.
  • Majolo F, Oliveira Paludo FJ, Ponzoni A, Graebin P, Dias FS, Alho CS. 2015. Effect of 593C>T GPx1 SNP alone and in synergy with 47C>T SOD2 SNP on the outcome of critically ill patients. Cytokine. 71(2):312–317.
  • Manczak M, Mao P, Calkins MJ, Cornea A, Reddy AP, Murphy MP, Szeto HH, Park B, Reddy PH. 2010. Mitochondria-targeted antioxidants protect against amyloid-beta toxicity in Alzheimer’s disease neurons. J Alzheimers Dis. 20(s2):S609–S631.
  • Manskikh VN, Gancharova OS, Nikiforova AI, Krasilshchikova MS, Shabalina IG, Egorov MV, Karger EM, Milanovsky GE, Galkin II, Skulachev VP, et al. 2015. Age-associated murine cardiac lesions are attenuated by the mitochondria-targeted antioxidant SkQ1. Histol Histopathol. 30(3):353–360.
  • Mansouri A, Tarhuni A, Larosche I, Reyl-Desmars F, Demeilliers C, Degoul F, Nahon P, Sutton A, Moreau R, Fromenty B, et al. 2010. MnSOD overexpression prevents liver mitochondrial DNA depletion after an alcohol binge but worsens this effect after prolonged alcohol consumption in mice. Dig Dis. 28(6):756–775.
  • Mao G, Kraus GA, Kim I, Spurlock ME, Bailey TB, Zhang Q, Beitz DC. 2010. A mitochondria-targeted vitamin E derivative decreases hepatic oxidative stress and inhibits fat deposition in mice. J Nutr. 140(8):1425–1431.
  • Mao P, Manczak M, Calkins MJ, Truong Q, Reddy TP, Reddy AP, Shirendeb U, Lo HH, Rabinovitch PS, Reddy PH. 2012. Mitochondria-targeted catalase reduces abnormal APP processing, amyloid beta production and BACE1 in a mouse model of Alzheimer’s disease: implications for neuroprotection and lifespan extension. Hum Mol Genet. 21(13):2973–2990.
  • Mao P, Manczak M, Shirendeb UP, Reddy PH. 2013. MitoQ, a mitochondria-targeted antioxidant, delays disease progression and alleviates pathogenesis in an experimental autoimmune encephalomyelitis mouse model of multiple sclerosis. Biochim Biophys Acta. 1832(12):2322–2331.
  • Mao C, Qiu LX, Zhan P, Xue K, Ding H, Du FB, Li J, Chen Q. 2010. MnSOD Val16Ala polymorphism and prostate cancer susceptibility: a meta-analysis involving 8,962 subjects. J Cancer Res Clin Oncol. 136(7):975–979.
  • Mapuskar KA, Wen H, Holanda DG, Rastogi P, Steinbach E, Han R, Coleman MC, Attanasio M, Riley DP, Spitz DR, et al. 2019. Persistent increase in mitochondrial superoxide mediates cisplatin-induced chronic kidney disease. Redox Biol. 20:98–106.
  • Maragos WF, Jakel R, Chesnut D, Pocernich CB, Butterfield DA, St Clair D, Cass WA. 2000. Methamphetamine toxicity is attenuated in mice that overexpress human manganese superoxide dismutase. Brain Res. 878(1–2):218–222.
  • Marei WFA, Van den Bosch L, Pintelon I, Mohey-Elsaeed O, Bols PEJ, Leroy J. 2019. Mitochondria-targeted therapy rescues development and quality of embryos derived from oocytes matured under oxidative stress conditions: a bovine in vitro model. Hum Reprod. 34(10):1984–1998.
  • Mari M, Bai J, Cederbaum AI. 2002. Adenovirus-mediated overexpression of catalase in the cytosolic or mitochondrial compartment protects against toxicity caused by glutathione depletion in HepG2 cells expressing CYP2E1. J Pharmacol Exp Ther. 301(1):111–118.
  • Mari M, Morales A, Colell A, Garcia-Ruiz C, Fernandez-Checa JC. 2009. Mitochondrial glutathione, a key survival antioxidant. Antioxid Redox Signaling. 11(11):2685–2700.
  • Marin-Royo G, Rodriguez C, Le Pape A, Jurado-Lopez R, Luaces M, Antequera A, Martinez-Gonzalez J, Souza-Neto FV, Nieto ML, Martinez-Martinez E, et al. 2019. The role of mitochondrial oxidative stress in the metabolic alterations in diet-induced obesity in rats. FASEB J. 33(11):12060–12072.
  • Marrotte EJ, Chen DD, Hakim JS, Chen AF. 2010. Manganese superoxide dismutase expression in endothelial progenitor cells accelerates wound healing in diabetic mice. J Clin Invest. 120(12):4207–4219.
  • Martin RC, Ahn J, Nowell SA, Hein DW, Doll MA, Martini BD, Ambrosone CB. 2006. Association between manganese superoxide dismutase promoter gene polymorphism and breast cancer survival. Breast Cancer Res. 8(4):R45.
  • Marullo R, Werner E, Degtyareva N, Moore B, Altavilla G, Ramalingam SS, Doetsch PW. 2013. Cisplatin induces a mitochondrial-ROS response that contributes to cytotoxicity depending on mitochondrial redox status and bioenergetic functions. PLoS One. 8(11):e81162.
  • Massaad CA, Amin SK, Hu L, Mei Y, Klann E, Pautler RG. 2010. Mitochondrial superoxide contributes to blood flow and axonal transport deficits in the Tg2576 mouse model of Alzheimer’s disease. PLoS One. 5(5):e10561.
  • Massaad CA, Washington TM, Pautler RG, Klann E. 2009. Overexpression of SOD-2 reduces hippocampal superoxide and prevents memory deficits in a mouse model of Alzheimer’s disease. Proc Nat Acad Sci USA. 106(32):13576–13581.
  • Matsushima S, Ide T, Yamato M, Matsusaka H, Hattori F, Ikeuchi M, Kubota T, Sunagawa K, Hasegawa Y, Kurihara T, et al. 2006. Overexpression of mitochondrial peroxiredoxin-3 prevents left ventricular remodeling and failure after myocardial infarction in mice. Circulation. 113(14):1779–1786.
  • Mattey DL, Hassell AB, Dawes PT, Jones PW, Yengi L, Alldersea J, Strange RC, Fryer AA. 2000. Influence of polymorphism in the manganese superoxide dismutase locus on disease outcome in rheumatoid arthritis: evidence for interaction with glutathione S-transferase genes. Arthritis Rheum. 43(4):859–864.
  • Mazumder S, De R, Sarkar S, Siddiqui AA, Saha SJ, Banerjee C, Iqbal MS, Nag S, Debsharma S, Bandyopadhyay U. 2016. Selective scavenging of intra-mitochondrial superoxide corrects diclofenac-induced mitochondrial dysfunction and gastric injury: a novel gastroprotective mechanism independent of gastric acid suppression. Biochem Pharmacol. 121:33–51.
  • McCarthy C, Kenny LC. 2016. Therapeutically targeting mitochondrial redox signalling alleviates endothelial dysfunction in preeclampsia. Sci Rep. 6:32683.
  • McClung JM, Deruisseau KC, Whidden MA, Van Remmen H, Richardson A, Song W, Vrabas IS, Powers SK. 2010. Overexpression of antioxidant enzymes in diaphragm muscle does not alter contraction-induced fatigue or recovery. Exp Physiol. 95(1):222–231.
  • McKnight AJ, Patterson CC, Mollsten A, Vance DR, Tarnow L, Maxwell AP. 2012. Review of genetic association in the SOD2 gene with chronic kidney disease: case-control studies and meta-analysis confirm association with diabetic nephropathy. Nephrol Rev. 4:e12.
  • McLachlan J, Beattie E, Murphy MP, Koh-Tan CH, Olson E, Beattie W, Dominiczak AF, Nicklin SA, Graham D. 2014. Combined therapeutic benefit of mitochondria-targeted antioxidant, MitoQ10, and angiotensin receptor blocker, losartan, on cardiovascular function. J Hypertens. 32(3):555–564.
  • McManus MJ, Murphy MP, Franklin JL. 2011. The mitochondria-targeted antioxidant MitoQ prevents loss of spatial memory retention and early neuropathology in a transgenic mouse model of Alzheimer’s disease. J Neurosci. 31(44):15703–15715.
  • Mehta SL, Lin Y, Chen W, Yu F, Cao L, He Q, Chan PH, Li PA. 2011. Manganese superoxide dismutase deficiency exacerbates ischemic brain damage under hyperglycemic conditions by altering autophagy. Transl Stroke Res. 2(1):42–50.
  • Melov S, Coskun P, Patel M, Tuinstra R, Cottrell B, Jun AS, Zastawny TH, Dizdaroglu M, Goodman SI, Huang TT, et al. 1999. Mitochondrial disease in superoxide dismutase 2 mutant mice. Proc Natl Acad Sci USA. 96(3):846–851.
  • Menon SG, Sarsour EH, Kalen AL, Venkataraman S, Hitchler MJ, Domann FE, Oberley LW, Goswami PC. 2007. Superoxide signaling mediates N-acetyl-L-cysteine-induced G1 arrest: regulatory role of cyclin D1 and manganese superoxide dismutase. Cancer Res. 67(13):6392–6399.
  • Mercer JR, Yu E, Figg N, Cheng KK, Prime TA, Griffin JL, Masoodi M, Vidal-Puig A, Murphy MP, Bennett MR. 2012. The mitochondria-targeted antioxidant MitoQ decreases features of the metabolic syndrome in ATM+/-/ApoE-/- mice. Free Radic Biol Med. 52(5):841–849.
  • Mezera V, Gerencser AA, Brand MD. 2017. Tunicamycin exposure triggers apopotosis by superoxide formation from site IIIQo of mitochondrial complex III [Abstract]. Free Radic Biol Med. 112(Supplement 1):172.
  • Mikhak B, Hunter DJ, Spiegelman D, Platz EA, Wu K, Erdman JW, Jr., Giovannucci E. 2008. Manganese superoxide dismutase (MnSOD) gene polymorphism, interactions with carotenoid levels and prostate cancer risk. Carcinogenesis. 29(12):2335–2340.
  • Millikan RC, Player J, de Cotret AR, Moorman P, Pittman G, Vannappagari V, Tse CK, Keku T. 2004. Manganese superoxide dismutase Ala-9Val polymorphism and risk of breast cancer in a population-based case-control study of African Americans and whites. Breast Cancer Res. 6(4):R264–R274.
  • Min K, Kwon OS, Smuder AJ, Wiggs MP, Sollanek KJ, Christou DD, Yoo JK, Hwang MH, Szeto HH, Kavazis AN, et al. 2015. Increased mitochondrial emission of reactive oxygen species and calpain activation are required for doxorubicin-induced cardiac and skeletal muscle myopathy. J Physiol (Lond). 593(8):2017–2036.
  • Min K, Smuder AJ, Kwon OS, Kavazis AN, Szeto HH, Powers SK. 2011. Mitochondrial-targeted antioxidants protect skeletal muscle against immobilization-induced muscle atrophy. J Appl Physiol. 111(5):1459–1466.
  • Miquel E, Cassina A, Martinez-Palma L, Souza JM, Bolatto C, Rodriguez-Bottero S, Logan A, Smith RA, Murphy MP, Barbeito L, et al. 2014. Neuroprotective effects of the mitochondria-targeted antioxidant MitoQ in a model of inherited amyotrophic lateral sclerosis. Free Radic Biol Med. 70:204–213.
  • Mistry Y, Poolman T, Williams B, Herbert KE. 2013. A role for mitochondrial oxidants in stress-induced premature senescence of human vascular smooth muscle cells. Redox Biol. 1:411–417.
  • Mitchell W, Ng EA, Tamucci JD, Boyd KJ, Sathappa M, Coscia A, Pan M, Han X, Eddy NA, May ER, et al. 2020. The mitochondria-targeted peptide SS-31 binds lipid bilayers and modulates surface electrostatics as a key component of its mechanism of action. J Biol Chem. 295(21):7452–7469.
  • Mitchell T, Rotaru D, Saba H, Smith RA, Murphy MP, MacMillan-Crow LA. 2011. The mitochondria-targeted antioxidant mitoquinone protects against cold storage injury of renal tubular cells and rat kidneys. J Pharmacol Exp Ther. 336(3):682–692.
  • Mitrunen K, Sillanpaa P, Kataja V, Eskelinen M, Kosma VM, Benhamou S, Uusitupa M, Hirvonen A. 2001. Association between manganese superoxide dismutase (MnSOD) gene polymorphism and breast cancer risk. Carcinogenesis. 22(5):827–829.
  • Miura S, Saitoh SI, Kokubun T, Owada T, Yamauchi H, Machii H, Takeishi Y. 2017. Mitochondrial-targeted antioxidant maintains blood flow, mitochondrial function, and redox balance in old mice following prolonged limb ischemia. Int J Mol Sci. 18(9):1897.
  • Mizuguchi Y, Chen J, Seshan SV, Poppas DP, Szeto HH, Felsen D. 2008. A novel cell-permeable antioxidant peptide decreases renal tubular apoptosis and damage in unilateral ureteral obstruction. Am J Physiol Renal Physiol. 295(5):F1545–F1553.
  • Mlakar SJ, Osredkar J, Prezelj J, Marc J. 2012. Antioxidant enzymes GSR, SOD1, SOD2, and CAT gene variants and bone mineral density values in postmenopausal women: a genetic association analysis. Menopause. 19(3):368–376.
  • Mohammedi K, Bellili-Munoz N, Driss F, Roussel R, Seta N, Fumeron F, Hadjadj S, Marre M, Velho G. 2014. Manganese superoxide dismutase (SOD2) polymorphisms, plasma advanced oxidation protein products (AOPP) concentration and risk of kidney complications in subjects with type 1 diabetes. PLoS One. 9(5):e96916.
  • Mohan SS, Ping XD, Harris FL, Ronda NJ, Brown LA, Gauthier TW. 2015. Fatty acid ethyl esters disrupt neonatal alveolar macrophage mitochondria and derange cellular functioning. Alcohol Clin Exp Res. 39(3):434–444.
  • Mollsten A, Jorsal A, Lajer M, Vionnet N, Tarnow L. 2009. The V16A polymorphism in SOD2 is associated with increased risk of diabetic nephropathy and cardiovascular disease in type 1 diabetes. Diabetologia. 52(12):2590–2593.
  • Mollsten A, Marklund SL, Wessman M, Svensson M, Forsblom C, Parkkonen M, Brismar K, Groop PH, Dahlquist G. 2007. A functional polymorphism in the manganese superoxide dismutase gene and diabetic nephropathy. Diabetes. 56(1):265–269.
  • Montalvo RN, Doerr V, Min K, Szeto HH, Smuder AJ. 2020. Doxorubicin-induced oxidative stress differentially regulates proteolytic signaling in cardiac and skeletal muscle. Am J Physiol Regul Integr Comp Physiol. 318(2):R227–R233.
  • Montano MA, Barrio Lera JP, Gottlieb MG, Schwanke CH, da Rocha MI, Manica-Cattani MF, dos Santos GF, da Cruz IB. 2009. Association between manganese superoxide dismutase (MnSOD) gene polymorphism and elderly obesity. Mol Cell Biochem. 328(1–2):33–40.
  • Montano MAE, Da Cruz IBM, Duarte MMMF, Krewer CdC, da Rocha MIdUM, Mânica-Cattani MF, Soares FAA, Rosa G, Maris AF, Battiston FG, et al. 2012. Inflammatory cytokines in vitro production are associated with Ala16Val superoxide dismutase gene polymorphism of peripheral blood mononuclear cells. Cytokine. 60(1):30–33.
  • Moshtaghi A, Vaziri H, Sariri R, Shaigan H. 2017. Polymorphism of MnSOD (Val16Ala) gene in pregnancies with blighted ovum: a case-control study. Int J Reprod Biomed. 15(8):503–508.
  • Mukhopadhyay P, Horvath B, Zsengeller Z, Zielonka J, Tanchian G, Holovac E, Kechrid M, Patel V, Stillman IE, Parikh SM, et al. 2012. Mitochondrial-targeted antioxidants represent a promising approach for prevention of cisplatin-induced nephropathy. Free Radic Biol Med. 52(2):497–506.
  • Mukhopadhyay P, Horváth B, Zsengellėr Z, Bátkai S, Cao Z, Kechrid M, Holovac E, Erdėlyi K, Tanchian G, Liaudet L, et al. 2012. Mitochondrial reactive oxygen species generation triggers inflammatory response and tissue injury associated with hepatic ischemia-reperfusion: therapeutic potential of mitochondrially targeted antioxidants. Free Radic Biol Med. 53(5):1123–1138.
  • Muliyil S, Narasimha M. 2014. Mitochondrial ROS regulates cytoskeletal and mitochondrial remodeling to tune cell and tissue dynamics in a model for wound healing. Dev Cell. 28(3):239–252.
  • Muller FL, Liu Y, Jernigan A, Borchelt D, Richardson A, Van Remmen H. 2008. MnSOD deficiency has a differential effect on disease progression in two different ALS mutant mouse models. Muscle Nerve. 38(3):1173–1183.
  • Muller FL, Liu Y, Van Remmen H. 2004. Complex III releases superoxide to both sides of the inner mitochondrial membrane. J Biol Chem. 279(47):49064–49073.
  • Murakami K, Kondo T, Kawase M, Li Y, Sato S, Chen SF, Chan PH. 1998. Mitochondrial susceptibility to oxidative stress exacerbates cerebral infarction that follows permanent focal cerebral ischemia in mutant mice with manganese superoxide dismutase deficiency. J Neurosci. 18(1):205–213.
  • Murphy MP. 2009. How mitochondria produce reactive oxygen species. Biochem J. 417(1):1–13.
  • Murphy MP. 2016. Understanding and preventing mitochondrial oxidative damage. Biochem Soc Trans. 44(5):1219–1226.
  • Murphy MP, Hartley RC. 2018. Mitochondria as a therapeutic target for common pathologies. Nat Rev Drug Discov. 17(12):865–886.
  • Murphy MP, Holmgren A, Larsson NG, Halliwell B, Chang CJ, Kalyanaraman B, Rhee SG, Thornalley PJ, Partridge L, Gems D, et al. 2011. Unraveling the biological roles of reactive oxygen species. Cell Metab. 13(4):361–366.
  • Nacarelli T, Azar A, Sell C. 2016. Mitochondrial stress induces cellular senescence in an mTORC1-dependent manner. Free Radic Biol Med. 95:133–154.
  • Nadatani Y, Huo X, Zhang X, Yu C, Cheng E, Zhang Q, Dunbar KB, Theiss A, Pham TH, Wang DH, et al. 2016. NOD-Like Receptor Protein 3 inflammasome priming and activation in Barrett’s epithelial cells. Cell Mol Gastroenterol Hepatol. 2(4):439–453.
  • Nagano T, Takeyama M. 2001. Enhancement of salivary secretion and neuropeptide (substance P, alpha-calcitonin gene-related peptide) levels in saliva by chronic anethole trithione treatment. J Pharm Pharmacol. 53(12):1697–1702.
  • Nahon P, Sutton A, Pessayre D, Rufat P, Degoul F, Ganne-Carrie N, Ziol M, Charnaux N, N’Kontchou G, Trinchet JC, et al. 2005. Genetic dimorphism in superoxide dismutase and susceptibility to alcoholic cirrhosis, hepatocellular carcinoma, and death. Clin Gastroenterol Hepatol. 3(3):292–298.
  • Nakagawa T, Kajiwara A, Saruwatari J, Hamamoto A, Kaku W, Oniki K, Mihara S, Ogata Y, Nakagawa K. 2013. The combination of mitochondrial low enzyme-activity aldehyde dehydrogenase 2 allele and superoxide dismutase 2 genotypes increases the risk of hypertension in relation to alcohol consumption. Pharmacogenet Genomics. 23(1):34–37.
  • Nakamura K, Chen CK, Sekine Y, Iwata Y, Anitha A, Loh el W, Takei N, Suzuki A, Kawai M, Takebayashi K, et al. 2006. Association analysis of SOD2 variants with methamphetamine psychosis in Japanese and Taiwanese populations. Hum Genet. 120(2):243–252.
  • Namikawa C, Shu-Ping Z, Vyselaar JR, Nozaki Y, Nemoto Y, Ono M, Akisawa N, Saibara T, Hiroi M, Enzan H, et al. 2004. Polymorphisms of microsomal triglyceride transfer protein gene and manganese superoxide dismutase gene in non-alcoholic steatohepatitis. J Hepatol. 40(5):781–786.
  • Naser NM, Babeker SE, Abdo M, Ismail AM, Altoum S, Bakheet KH. 2019. Association of MnSOD-47C/TGENE polymorphism (rs 4880-47C/T) in Sudanese patient with diabetic retinopathy. Asian J Med Health. 15:1–9.
  • Nautiyal M, Shaltout HA, Chappell MC, Diz DI. 2019. Comparison of Candesartan and Angiotensin-(1-7) combination to Mito-TEMPO treatment for normalizing blood pressure and sympathovagal balance in (mREN2)27 rats. J Cardiovasc Pharmacol. 73(3):143–148.
  • Nazarewicz RR, Dikalova AE, Bikineyeva A, Dikalov SI. 2013. Nox2 as a potential target of mitochondrial superoxide and its role in endothelial oxidative stress. Am J Physiol Heart Circ Physiol. 305(8):H1131–1140.
  • Nazarewicz RR, Dikalova A, Bikineyeva A, Ivanov S, Kirilyuk IA, Grigor’ev IA, Dikalov SI. 2013. Does scavenging of mitochondrial superoxide attenuate cancer prosurvival signaling pathways? Antioxid Redox Signal. 19(4):344–349.
  • Nemoto Y, Yamamoto T, Takada S, Matsui Y, Obinata M. 1990. Antisense RNA of the latent period gene (MER5) inhibits the differentiation of murine erythroleukemia cells. Gene. 91(2):261–265.
  • Neroev VV, Archipova MM, Bakeeva LE, Fursova A, Grigorian EN, Grishanova AY, Iomdina EN, Ivashchenko ZN, Katargina LA, Khoroshilova-Maslova IP, et al. 2008. Mitochondria-targeted plastoquinone derivatives as tools to interrupt execution of the aging program. 4. Age-related eye disease. SkQ1 returns vision to blind animals. Biochemistry Mosc. 73(12):1317–1328.
  • Neuzil J, Widen C, Gellert N, Swettenham E, Zobalova R, Dong LF, Wang XF, Lidebjer C, Dalen H, Headrick JP, et al. 2007. Mitochondria transmit apoptosis signalling in cardiomyocyte-like cells and isolated hearts exposed to experimental ischemia-reperfusion injury. Redox Rep. 12(3):148–162.
  • Ni R, Cao T, Xiong S, Ma J, Fan GC, Lacefield JC, Lu Y, Le Tissier S, Peng T. 2016. Therapeutic inhibition of mitochondrial reactive oxygen species with mito-TEMPO reduces diabetic cardiomyopathy. Free Radic Biol Med. 90:12–23.
  • Nishikawa T, Edelstein D, Du XL, Yamagishi S, Matsumura T, Kaneda Y, Yorek MA, Beebe D, Oates PJ, Hammes HP, et al. 2000. Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature. 404(6779):787–790.
  • Niu Y, Epperly MW, Shen H, Smith T, Wang H, Greenberger JS. 2008. Intraesophageal MnSOD-plasmid liposome enhances engraftment and self-renewal of bone marrow derived progenitors of esophageal squamous epithelium. Gene Ther. 15(5):347–356.
  • Niu Y, Wang H, Wiktor-Brown D, Rugo R, Shen H, Huq MS, Engelward B, Epperly M, Greenberger JS. 2010. Irradiated esophageal cells are protected from radiation-induced recombination by MnSOD gene therapy. Radiat Res. 173(4):453–461.
  • Nojiri H, Shimizu T, Funakoshi M, Yamaguchi O, Zhou H, Kawakami S, Ohta Y, Sami M, Tachibana T, Ishikawa H, et al. 2006. Oxidative stress causes heart failure with impaired mitochondrial respiration. J Biol Chem. 281(44):33789–33801.
  • Nolan LS, Cadge BA, Gomez-Dorado M, Dawson SJ. 2013. A functional and genetic analysis of SOD2 promoter variants and their contribution to age-related hearing loss. Mech Ageing Dev. 134(7–8):298–306.
  • Nomiyama T, Tanaka Y, Piao L, Nagasaka K, Sakai K, Ogihara T, Nakajima K, Watada H, Kawamori R. 2003. The polymorphism of manganese superoxide dismutase is associated with diabetic nephropathy in Japanese type 2 diabetic patients. J Hum Genet. 48(3):138–141.
  • Nonn L, Berggren M, Powis G. 2003. Increased expression of mitochondrial peroxiredoxin-3 (thioredoxin peroxidase-2) protects cancer cells against hypoxia and drug-induced hydrogen peroxide-dependent apoptosis. Mol Cancer Res. 1(9):682–689.
  • Nussbaumer M, Asara JM, Teplytska L, Murphy MP, Logan A, Turck CW, Filiou MD. 2016. Selective mitochondrial targeting exerts anxiolytic effects in vivo. Neuropsychopharmacology. 41(7):1751–1758.
  • Nuzzo AM, Camm EJ, Sferruzzi-Perri AN, Ashmore TJ, Yung HW, Cindrova-Davies T, Spiroski AM, Sutherland MR, Logan A, Austin-Williams S, et al. 2018. Placental adaptation to early-onset hypoxic pregnancy and mitochondria-targeted antioxidant therapy in a rodent model. Am J Pathol. 188(12):2704–2716.
  • Oberkampf M, Guillerey C, Mouries J, Rosenbaum P, Fayolle C, Bobard A, Savina A, Ogier-Denis E, Enninga J, Amigorena S, et al. 2018. Mitochondrial reactive oxygen species regulate the induction of CD8(+) T cells by plasmacytoid dendritic cells. Nat Commun. 9(1):2241.
  • Obukhova LA, Skulachev VP, Kolosova NG. 2009. Mitochondria-targeted antioxidant SkQ1 inhibits age-dependent involution of the thymus in normal and senescence-prone rats. Aging (Albany NY). 1(4):389–401.
  • Ohashi M, Runge MS, Faraci FM, Heistad DD. 2006. MnSOD deficiency increases endothelial dysfunction in ApoE-deficient mice. Arterioscler Thromb Vasc Biol. 26(10):2331–2336.
  • Ojano-Dirain CP, Antonelli PJ. 2012. Prevention of gentamicin-induced apoptosis with the mitochondria-targeted antioxidant mitoquinone. Laryngoscope. 122(11):2543–2548.
  • Ojano-Dirain CP, Antonelli PJ, Le Prell CG. 2014. Mitochondria-targeted antioxidant MitoQ reduces gentamicin-induced ototoxicity. Otol Neurotol. 35(3):533–539.
  • Olesen MA, Torres AK, Jara C, Murphy MP, Tapia-Rojas C. 2020. Premature synaptic mitochondrial dysfunction in the hippocampus during aging contributes to memory loss. Redox Biol. 34:101558.
  • Olgar Y, Billur D, Tuncay E, Turan B. 2020. MitoTEMPO provides an antiarrhythmic effect in aged-rats through attenuation of mitochondrial reactive oxygen species. Exp Gerontol. 136:110961.
  • Olgar Y, Degirmenci S, Durak A, Billur D, Can B, Kayki-Mutlu G, Arioglu-Inan EE, Turan B. 2018. Aging related functional and structural changes in the heart and aorta: MitoTEMPO improves aged-cardiovascular performance. Exp Gerontol. 110:172–181.
  • Olsen RH, Johnson LA, Zuloaga DG, Limoli CL, Raber J. 2013. Enhanced hippocampus-dependent memory and reduced anxiety in mice over-expressing human catalase in mitochondria. J Neurochem. 125(2):303–313.
  • Olson DH, Burrill JS, Kuzmicic J, Hahn WS, Park JM, Kim DH, Bernlohr DA. 2017. Down regulation of Peroxiredoxin-3 in 3T3-L1 adipocytes leads to oxidation of Rictor in the mammalian-target of rapamycin complex 2 (mTORC2). Biochem Biophys Res Commun. 493(3):1311–1317.
  • Orhan N, Kucukali CI, Cakir U, Seker N, Aydin M. 2012. Genetic variants in nuclear-encoded mitochondrial proteins are associated with oxidative stress in obsessive compulsive disorders. J Psychiatr Res. 46(2):212–218.
  • Orr AL, Ashok D, Sarantos MR, Shi T, Hughes RE, Brand MD. 2013. Inhibitors of ROS production by the ubiquinone-binding site of mitochondrial complex I identified by chemical screening. Free Radic Biol Med. 65:1047–1059.
  • Orr AL, Vargas L, Turk CN, Baaten JE, Matzen JT, Dardov VJ, Attle SJ, Li J, Quackenbush DC, Goncalves RLS, et al. 2015. Suppressors of superoxide production from mitochondrial complex III. Nat Chem Biol. 11(11):834–836.
  • Ough M, Lewis A, Zhang Y, Hinkhouse MM, Ritchie JM, Oberley LW, Cullen JJ. 2004. Inhibition of cell growth by overexpression of manganese superoxide dismutase (MnSOD) in human pancreatic carcinoma. Free Radic Res. 38(11):1223–1233.
  • Owada T, Yamauchi H, Saitoh SI, Miura S, Machii H, Takeishi Y. 2017. Resolution of mitochondrial oxidant stress improves aged-cardiovascular performance. Coron Artery Dis. 28(1):33–43.
  • Ozeki M, Tamae D, Hou DX, Wang T, Lebon T, Spitz DR, Li JJ. 2004. Response of cyclin B1 to ionizing radiation: regulation by NF-kappaB and mitochondrial antioxidant enzyme MnSOD. Anticancer Res. 24(5A):2657–2663.
  • Pacini G, Mari A, Fouqueray P, Bolze S, Roden M. 2015. Imeglimin increases glucose-dependent insulin secretion and improves beta-cell function in patients with type 2 diabetes. Diabetes Obes Metab. 17(6):541–545.
  • Pae CU, Kim TS, Patkar AA, Kim JJ, Lee CU, Lee SJ, Jun TY, Lee C, Paik IH. 2007. Manganese superoxide dismutase (MnSOD: Ala-9Val) gene polymorphism may not be associated with schizophrenia and tardive dyskinesia. Psychiatry Res. 153(1):77–81.
  • Pae CU, Yoon SJ, Patkar A, Kim JJ, Jun TY, Lee C, Paik IH. 2006. Manganese superoxide dismutase (MnSOD: Ala-9Val) gene polymorphism and mood disorders: a preliminary study. Prog Neuropsychopharmacol Biol Psychiatry. 30(7):1326–1329.
  • Paglialunga S, Ludzki A, Root-McCaig J, Holloway GP. 2015. In adipose tissue, increased mitochondrial emission of reactive oxygen species is important for short-term high-fat diet-induced insulin resistance in mice. Diabetologia. 58(5):1071–1080.
  • Paglialunga S, van Bree B, Bosma M, Valdecantos MP, Amengual-Cladera E, Jorgensen JA, van Beurden D, den Hartog GJM, Ouwens DM, Briede JJ, et al. 2012. Targeting of mitochondrial reactive oxygen species production does not avert lipid-induced insulin resistance in muscle tissue from mice. Diabetologia. 55(10):2759–2768.
  • Pak O, Scheibe S, Esfandiary A, Gierhardt M, Sydykov A, Logan A, Fysikopoulos A, Veit F, Hecker M, Kroschel F, et al. 2018. Impact of the mitochondria-targeted antioxidant MitoQ on hypoxia-induced pulmonary hypertension. Eur Respir J. 51(3):1701024.
  • Palmirotta R, Barbanti P, De Marchis ML, Egeo G, Aurilia C, Fofi L, Ialongo C, Valente MG, Ferroni P, Della-Morte D, et al. 2015. Is SOD2 Ala16Val polymorphism associated with migraine with aura phenotype? Antioxid Redox Signal. 22(3):275–279.
  • Paludo F. J d O, Picanço JB, Fallavena PRV, Fraga L. d R, Graebin P, Nóbrega O. d T, Dias FS, Alho CS. 2013. Higher frequency of septic shock in septic patients with the 47C allele (rs4880) of the SOD2 gene. Gene. 517(1):106–111.
  • Parajuli N, Campbell LH, Marine A, Brockbank KG, Macmillan-Crow LA. 2012. MitoQ blunts mitochondrial and renal damage during cold preservation of porcine kidneys. PLoS One. 7(11):e48590.
  • Parihar VK, Allen BD, Tran KK, Chmielewski NN, Craver BM, Martirosian V, Morganti JM, Rosi S, Vlkolinsky R, Acharya MM, et al. 2015. Targeted overexpression of mitochondrial catalase prevents radiation-induced cognitive dysfunction. Antioxid Redox Signal. 22(1):78–91.
  • Park E, Chung SW. 2019. ROS-mediated autophagy increases intracellular iron levels and ferroptosis by ferritin and transferrin receptor regulation. Cell Death Dis. 10(11):822.
  • Park JH, Ku HJ, Lee JH, Park JW. 2018. IDH2 deficiency accelerates skin pigmentation in mice via enhancing melanogenesis. Redox Biol. 17:16–24.
  • Park J, Min JS, Kim B, Chae UB, Yun JW, Choi MS, Kong IK, Chang KT, Lee DS. 2015. Mitochondrial ROS govern the LPS-induced pro-inflammatory response in microglia cells by regulating MAPK and NF-kappaB pathways. Neurosci Lett. 584:191–196.
  • Park JB, Nagar H, Choi S, Jung SB, Kim HW, Kang SK, Lee JW, Lee JH, Park JW, Irani K, et al. 2016. IDH2 deficiency impairs mitochondrial function in endothelial cells and endothelium-dependent vasomotor function. Free Radic Biol Med. 94:36–46.
  • Parlaktas BS, Atilgan D, Gencten Y, Benli I, Ozyurt H, Uluocak N, Erdemir F. 2015. A pilot study of the association of manganese superoxide dismutase and glutathione peroxidase 1 single gene polymorphisms with prostate cancer and serum prostate specific antigen levels. Arch Med Sci. 11(5):994–1000.
  • Pascotini MET, Flores DAE, Kegler MA, Konzen MV, Fornari MAL, Arend MJ, Gabbi MP, Gobo MLA, Bochi DGV, Prado D, et al. 2018. Brain-derived neurotrophic factor levels are lower in chronic stroke patients: a relation with manganese-dependent superoxide dismutase ALA16VAL single nucleotide polymorphism through tumor necrosis factor-alpha and caspases pathways. J Stroke Cerebrovasc Dis. 27(11):3020–3029.
  • Patil NK, Parajuli N, MacMillan-Crow LA, Mayeux PR. 2014. Inactivation of renal mitochondrial respiratory complexes and manganese superoxide dismutase during sepsis: mitochondria-targeted antioxidant mitigates injury. Am J Physiol Renal Physiol. 306(7):F734–F743.
  • Paul A, Belton A, Nag S, Martin I, Grotewiel MS, Duttaroy A. 2007. Reduced mitochondrial SOD displays mortality characteristics reminiscent of natural aging. Mech Ageing Dev. 128(11–12):706–716.
  • Pehar M, Beeson G, Beeson CC, Johnson JA, Vargas MR. 2014. Mitochondria-targeted catalase reverts the neurotoxicity of hSOD1G(9)(3)A astrocytes without extending the survival of ALS-linked mutant hSOD1 mice. PLoS One. 9(7):e103438.
  • Perez VI, Van Remmen H, Bokov A, Epstein CJ, Vijg J, Richardson A. 2009. The overexpression of major antioxidant enzymes does not extend the lifespan of mice. Aging Cell. 8(1):73–75.
  • Perier C, Bove J, Dehay B, Jackson-Lewis V, Rabinovitch PS, Przedborski S, Vila M. 2010. Apoptosis-inducing factor deficiency sensitizes dopaminergic neurons to parkinsonian neurotoxins. Ann Neurol. 68(2):184–192.
  • Perry RJ, Cardone RL, Petersen MC, Zhang D, Fouqueray P, Hallakou-Bozec S, Bolze S, Shulman GI, Petersen KF, Kibbey RG. 2016. Imeglimin lowers glucose primarily by amplifying glucose-stimulated insulin secretion in high-fat-fed rodents. Am J Physiol Endocrinol Metab. 311(2):E461–E470.
  • Peskin AV, Cox AG, Nagy P, Morgan PE, Hampton MB, Davies MJ, Winterbourn CC. 2010. Removal of amino acid, peptide and protein hydroperoxides by reaction with peroxiredoxins 2 and 3. Biochem J. 432(2):313–321.
  • Petri S, Kiaei M, Damiano M, Hiller A, Wille E, Manfredi G, Calingasan NY, Szeto HH, Beal MF. 2006. Cell-permeable peptide antioxidants as a novel therapeutic approach in a mouse model of amyotrophic lateral sclerosis. J Neurochem. 98(4):1141–1148.
  • Petrov A, Perekhvatova N, Skulachev M, Stein L, Ousler G. 2016. SkQ1 ophthalmic solution for dry eye treatment: results of a phase 2 safety and efficacy clinical study in the environment and during challenge in the controlled adverse environment model. Adv Ther. 33(1):96–115.
  • Petrovic MG, Cilensek I, Petrovic D. 2008. Manganese superoxide dismutase gene polymorphism (V16A) is associated with diabetic retinopathy in Slovene (Caucasians) type 2 diabetes patients. Dis Markers. 24(1):59–64.
  • Phillips TJ, Scott H, Menassa DA, Bignell AL, Sood A, Morton JS, Akagi T, Azuma K, Rogers MF, Gilmore CE, et al. 2017. Treating the placenta to prevent adverse effects of gestational hypoxia on fetal brain development. Sci Rep. 7(1):9079.
  • Piao L, Fang YH, Hamanaka RB, Mutlu GM, Dezfulian C, Archer SL, Sharp WW. 2020. Suppression of superoxide-hydrogen peroxide production at site IQ of mitochondrial complex I attenuates myocardial stunning and improves postcardiac arrest outcomes. Crit Care Med. 48(2):e133–e140.
  • Pietras T, Szemraj J, Witusik A, Hołub M, Panek M, Wujcik R, Górski P. 2010. The sequence polymorphism of MnSOD gene in subjects with respiratory insufficiency in COPD. Med Sci Monit. 16(9):CR427–CR432.
  • Pietras T, Witusik A, Panek M, Gałecki P, Szemraj J, Górski P. 2010. Anxiety, depression and polymorphism of the gene encoding superoxide dismutase in patients with chronic obstructive pulmonary disease. Pol Merkur Lekarski. 29(171):165–168.
  • Piippo N, Korhonen E, Hytti M, Kinnunen K, Kaarniranta K, Kauppinen A. 2018. Oxidative stress is the principal contributor to inflammasome activation in retinal pigment epithelium cells with defunct proteasomes and autophagy. Cell Physiol Biochem. 49(1):359–367.
  • Pinho BR, Duarte AI, Canas PM, Moreira PI, Murphy MP, Oliveira JMA. 2020. The interplay between redox signalling and proteostasis in neurodegeneration: In vivo effects of a mitochondria-targeted antioxidant in Huntington’s disease mice. Free Radic Biol Med. 146:372–382.
  • Pletjushkina OY, Fetisova EK, Lyamzaev KG, Ivanova OY, Domnina LV, Vyssokikh MY, Pustovidko AV, Alexeevski AV, Alexeevski DA, Vasiliev JM, et al. 2006. Hydrogen peroxide produced inside mitochondria takes part in cell-to-cell transmission of apoptotic signal. Biochem Mosc. 71(1):60–67.
  • Plotnikov EY, Chupyrkina AA, Jankauskas SS, Pevzner IB, Silachev DN, Skulachev VP, Zorov DB. 2011. Mechanisms of nephroprotective effect of mitochondria-targeted antioxidants under rhabdomyolysis and ischemia/reperfusion. Biochim Biophys Acta. 1812(1):77–86.
  • Plotnikov EY, Silachev DN, Chupyrkina AA, Danshina MI, Jankauskas SS, Morosanova MA, Stelmashook EV, Vasileva AK, Goryacheva ES, Pirogov YA, et al. 2010. New-generation Skulachev ions exhibiting nephroprotective and neuroprotective properties. Biochemistry Mosc. 75(2):145–150.
  • Pociot F, Lorenzen T, Nerup J. 1993. A manganese superoxide dismutase (SOD2) gene polymorphism in insulin-dependent diabetes mellitus. Dis Markers. 11(5–6):267–274.
  • Poggi C, Giusti B, Vestri A, Pasquini E, Abbate R, Dani C. 2012. Genetic polymorphisms of antioxidant enzymes in preterm infants. J Matern Fetal Neonatal Med. 25(Suppl 4):131–134.
  • Pohl K, Wall S, Locy M, Tipple T. 2018. Suppression of site IQ electron leak attenuates lung epithelial cell proliferation. Free Radical Biol Med. 128(Suppl. 1):S36.
  • Porporato PE, Payen VL, Perez-Escuredo J, De Saedeleer CJ, Danhier P, Copetti T, Dhup S, Tardy M, Vazeille T, Bouzin C, et al. 2014. A mitochondrial switch promotes tumor metastasis. Cell Rep. 8(3):754–766.
  • Pournourali M, Tarang A, Haghighi SF, Yousefi M, Bahadori MH. 2016. Polymorphism variant of MnSOD A16V and risk of female infertility in northern Iran. Taiwan J Obstet Gynecol. 55(6):801–803.
  • Pouzaud F, Christen MO, Warnet JM, Rat P. 2004. Anethole dithiolethione: an antioxidant agent against tenotoxicity induced by fluoroquinolones. Pathol Biol (Paris). 52(6):308–313.
  • Powell CS, Jackson RM. 2003. Mitochondrial complex I, aconitase, and succinate dehydrogenase during hypoxia-reoxygenation: modulation of enzyme activities by MnSOD. Am J Physiol Lung Cell Mol Physiol. 285(1):L189–L198.
  • Powell RD, Swet JH, Kennedy KL, Huynh TT, Murphy MP, McKillop IH, Evans SL. 2015. MitoQ modulates oxidative stress and decreases inflammation following hemorrhage. J Trauma Acute Care Surg. 78(3):573–579.
  • Powers SK, Hudson MB, Nelson WB, Talbert EE, Min K, Szeto HH, Kavazis AN, Smuder AJ. 2011. Mitochondria-targeted antioxidants protect against mechanical ventilation-induced diaphragm weakness. Crit Care Med. 39(7):1749–1759.
  • Prauchner CA. 2017. Oxidative stress in sepsis: pathophysiological implications justifying antioxidant co-therapy. Burns. 43(3):471–485.
  • Pung YF, Rocic P, Murphy MP, Smith RA, Hafemeister J, Ohanyan V, Guarini G, Yin L, Chilian WM. 2012. Resolution of mitochondrial oxidative stress rescues coronary collateral growth in Zucker obese fatty rats. Arterioscler Thromb Vasc Biol. 32(2):325–334.
  • Qi X, Lewin AS, Hauswirth WW, Guy J. 2003. Optic neuropathy induced by reductions in mitochondrial superoxide dismutase. Invest Ophthalmol Vis Sci. 44(3):1088–1096.
  • Qi X, Lewin AS, Sun L, Hauswirth WW, Guy J. 2004. SOD2 gene transfer protects against optic neuropathy induced by deficiency of complex I. Ann Neurol. 56(2):182–191.
  • Qi X, Sun L, Hauswirth WW, Lewin AS, Guy J. 2007. Use of mitochondrial antioxidant defenses for rescue of cells with a Leber hereditary optic neuropathy-causing mutation. Arch Ophthalmol. 125(2):268–272.
  • Quinlan CL, Gerencser AA, Treberg JR, Brand MD. 2011. The mechanism of superoxide production by the antimycin-inhibited mitochondrial Q-cycle. J Biol Chem. 286(36):31361–31372.
  • Quinlan CL, Goncalves RL, Hey-Mogensen M, Yadava N, Bunik VI, Brand MD. 2014. The 2-oxoacid dehydrogenase complexes in mitochondria can produce superoxide/hydrogen peroxide at much higher rates than complex I. J Biol Chem. 289(12):8312–8325.
  • Quinlan CL, Orr AL, Perevoshchikova IV, Treberg JR, Ackrell BA, Brand MD. 2012. Mitochondrial complex II can generate reactive oxygen species at high rates in both the forward and reverse reactions. J Biol Chem. 287(32):27255–27264.
  • Quinlan CL, Perevoschikova IV, Goncalves RL, Hey-Mogensen M, Brand MD. 2013. The determination and analysis of site-specific rates of mitochondrial reactive oxygen species production. Meth Enzymol. 526:189–217.
  • Quinlan CL, Treberg JR, Perevoshchikova IV, Orr AL, Brand MD. 2012. Native rates of superoxide production from multiple sites in isolated mitochondria measured using endogenous reporters. Free Radic Biol Med. 53(9):1807–1817.
  • Rademann P, Weidinger A, Drechsler S, Meszaros A, Zipperle J, Jafarmadar M, Dumitrescu S, Hacobian A, Ungelenk L, Rostel F, et al. 2017. Mitochondria-targeted antioxidants SkQ1 and MitoTEMPO failed to exert a long-term beneficial effect in murine polymicrobial sepsis. Oxid Med Cell Longev. 2017:6412682.
  • Raha S, Robinson BH. 2000. Mitochondria, oxygen free radicals, disease and ageing. Trends Biochem Sci. 25(10):502–508.
  • Raimundo N, Song L, Shutt TE, McKay SE, Cotney J, Guan MX, Gilliland TC, Hohuan D, Santos-Sacchi J, Shadel GS. 2012. Mitochondrial stress engages E2F1 apoptotic signaling to cause deafness. Cell. 148(4):716–726.
  • Raineri I, Carlson EJ, Gacayan R, Carra S, Oberley TD, Huang TT, Epstein CJ. 2001. Strain-dependent high-level expression of a transgene for manganese superoxide dismutase is associated with growth retardation and decreased fertility. Free Radic Biol Med. 31(8):1018–1030.
  • Rajić V, Aplenc R, Debeljak M, Prestor VV, Karas-Kuželicki N, MlinariČ-RašČan I, Jazbec J. 2009. Influence of the polymorphism in candidate genes on late cardiac damage in patients treated due to acute leukemia in childhood. Leuk Lymphoma. 50(10):1693–1698.
  • Ramachandran A, Lebofsky M, Weinman SA, Jaeschke H. 2011. The impact of partial manganese superoxide dismutase (SOD2)-deficiency on mitochondrial oxidant stress, DNA fragmentation and liver injury during acetaminophen hepatotoxicity. Toxicol Appl Pharmacol. 251(3):226–233.
  • Ramos GB, Salomao H, Francio AS, Fava VM, Werneck RI, Mira MT. 2016. Association analysis suggests SOD2 as a newly identified candidate gene associated with leprosy susceptibility. J Infect Dis. 214(3):475–478.
  • Ramprasath T, Murugan PS, Kalaiarasan E, Gomathi P, Rathinavel A, Selvam GS. 2012. Genetic association of Glutathione peroxidase-1 (GPx-1) and NAD(P)H:Quinone Oxidoreductase 1(NQO1) variants and their association of CAD in patients with type-2 diabetes. Mol Cell Biochem. 361(1-2):143–150.
  • Rao VR, Lautz JD, Kaja S, Foecking EM, Lukacs E, Stubbs EB. Jr. 2019. Mitochondrial-targeted antioxidants attenuate TGF-beta2 signaling in human trabecular meshwork cells. Invest Ophthalmol Vis Sci. 60(10):3613–3624.
  • Ratnasinghe D, Tangrea JA, Andersen MR, Barrett MJ, Virtamo J, Taylor PR, Albanes D. 2000. Glutathione peroxidase codon 198 polymorphism variant increases lung cancer risk. Cancer Res. 60(22):6381–6383.
  • Ray PD, Huang BW, Tsuji Y. 2012. Reactive oxygen species (ROS) homeostasis and redox regulation in cellular signaling. Cell Signal. 24(5):981–990.
  • Reddy VN, Kasahara E, Hiraoka M, Lin LR, Ho YS. 2004. Effects of variation in superoxide dismutases (SOD) on oxidative stress and apoptosis in lens epithelium. Exp Eye Res. 79(6):859–868.
  • Reddy TP, Manczak M, Calkins MJ, Mao P, Reddy AP, Shirendeb U, Park B, Reddy PH. 2011. Toxicity of neurons treated with herbicides and neuroprotection by mitochondria-targeted antioxidant SS31. Int J Environ Res Public Health. 8(1):203–221.
  • Reddy PH, Manczak M, Kandimalla R. 2017. Mitochondria-targeted small molecule SS31: a potential candidate for the treatment of Alzheimer’s disease. Hum Mol Genet. 26(8):1483–1496.
  • Reddy PH, Manczak M, Yin X, Reddy AP. 2018. Synergistic protective effects of mitochondrial division inhibitor 1 and mitochondria-targeted small peptide SS31 in Alzheimer’s disease. J Alzheimers Dis. 62(4):1549–1565.
  • Reddy BS, Rao CV, Rivenson A, Kelloff G. 1993. Chemoprevention of colon carcinogenesis by organosulfur compounds. Cancer Res. 53(15):3493–3498.
  • Rego I, Fernandez-Moreno M, Fernandez-Lopez C, Gomez-Reino JJ, Gonzalez A, Arenas J, Blanco FJ. 2010. Role of European mitochondrial DNA haplogroups in the prevalence of hip osteoarthritis in Galicia, Northern Spain. Ann Rheum Dis. 69(1):210–213.
  • Rego-Perez I, Fernandez-Moreno M, Fernandez-Lopez C, Arenas J, Blanco FJ. 2008. Mitochondrial DNA haplogroups: role in the prevalence and severity of knee osteoarthritis. Arthritis Rheum. 58(8):2387–2396.
  • Rehman H, Liu Q, Krishnasamy Y, Shi Z, Ramshesh VK, Haque K, Schnellmann RG, Murphy MP, Lemasters JJ, Rockey DC, et al. 2016. The mitochondria-targeted antioxidant MitoQ attenuates liver fibrosis in mice. Int J Physiol Pathophysiol Pharmacol. 8(1):14–27.
  • Reynolds TH, Dalton A, Calzini L, Tuluca A, Hoyte D, Ives SJ. 2019. The impact of age and sex on body composition and glucose sensitivity in C57BL/6J mice. Physiol Rep. 7(3):e13995.
  • Ribeiro Junior RF, Dabkowski ER, Shekar KC, O Connell KA, Hecker PA, Murphy MP. 2018. MitoQ improves mitochondrial dysfunction in heart failure induced by pressure overload. Free Radic Biol Med. 117:18–29.
  • Richters L, Lange N, Renner R, Treiber N, Ghanem A, Tiemann K, Scharffetter-Kochanek K, Bloch W, Brixius K. 2011. Exercise-induced adaptations of cardiac redox homeostasis and remodeling in heterozygous SOD2-knockout mice. J Appl Physiol (1985). 111(5):1431–1440.
  • Ridnour LA, Oberley TD, Oberley LW. 2004. Tumor suppressive effects of MnSOD overexpression may involve imbalance in peroxide generation versus peroxide removal. Antioxid Redox Signal. 6(3):501–512.
  • Righi V, Constantinou C, Mintzopoulos D, Khan N, Mupparaju SP, Rahme LG, Swartz HM, Szeto HH, Tompkins RG, Tzika AA. 2013. Mitochondria-targeted antioxidant promotes recovery of skeletal muscle mitochondrial function after burn trauma assessed by in vivo 31P nuclear magnetic resonance and electron paramagnetic resonance spectroscopy. FASEB J. 27(6):2521–2530.
  • Rigoulet M, Yoboue ED, Devin A. 2011. Mitochondrial ROS generation and its regulation: mechanisms involved in H2O2 signaling. Antioxid Redox Signaling. 14(3):459–468.
  • Rivera-Barahona A, Alonso-Barroso E, Perez B, Murphy MP, Richard E, Desviat LR. 2017. Treatment with antioxidants ameliorates oxidative damage in a mouse model of propionic acidemia. Mol Genet Metab. 122(1–2):43–50.
  • Robb EL, Hall AR, Prime TA, Eaton S, Szibor M, Viscomi C, James AM, Murphy MP. 2018. Control of mitochondrial superoxide production by reverse electron transport at complex I. J Biol Chem. 293(25):9869–9879.
  • Robb EL, Gawel JM, Aksentijević D, Cochemé HM, Stewart TS, Shchepinova MM, Qiang H, Prime TA, Bright TP, James AM, et al. 2015. Selective superoxide generation within mitochondria by the targeted redox cycler MitoParaquat. Free Radic Biol Med. 89:883–894.
  • Rocha VCJ, França LS, de Araújo CF, Ng AM, de Andrade CM, Andrade AC, Santos Ede S, Borges-Silva Mda C, Macambira SG, Noronha-Dutra AA, et al. 2016. Protective effects of mito-TEMPO against doxorubicin cardiotoxicity in mice. Cancer Chemother Pharmacol. 77(3):659–662.
  • Rodrı́guez AM, Carrico PM, Mazurkiewicz JE, Meléndez JA. 2000. Mitochondrial or cytosolic catalase reverses the MnSOD-dependent inhibition of proliferation by enhancing respiratory chain activity, net ATP production, and decreasing the steady state levels of H(2)O(2). Free Radic Biol Med. 29(9):801–813.
  • Rodriguez-Iturbe B, Sepassi L, Quiroz Y, Ni Z, Wallace DC, Vaziri ND. 2007. Association of mitochondrial SOD deficiency with salt-sensitive hypertension and accelerated renal senescence. J Appl Physiol (1985). 102(1):255–260.
  • Rossman MJ, Santos-Parker JR, Steward CAC, Bispham NZ, Cuevas LM, Rosenberg HL, Woodward KA, Chonchol M, Gioscia-Ryan RA, Murphy MP, et al. 2018. Chronic supplementation with a mitochondrial antioxidant (MitoQ) improves vascular function in healthy older adults. Hypertension. 71(6):1056–1063.
  • Roy N, Dasgupta D, Mukhopadhyay I, Chatterjee A, Das K, Bhowmik P, Das S, Basu P, Santra AK, Datta S, et al. 2016. Genetic association and gene-gene interaction reveal genetic variations in ADH1B, GSTM1 and MnSOD independently confer risk to alcoholic liver diseases in India. PLoS One. 11(3):e0149843.
  • Ruiz-Sanz JI, Aurrekoetxea I, Matorras R, Ruiz-Larrea MB. 2011. Ala16Val SOD2 polymorphism is associated with higher pregnancy rates in in vitro fertilization cycles. Fertil Steril. 95(5):1601–1605.
  • Ryan TE, Schmidt CA, Green TD, Spangenburg EE, Neufer PD, McClung JM. 2016. Targeted expression of catalase to mitochondria protects against ischemic myopathy in high-fat diet-fed mice. Diabetes. 65(9):2553–2568.
  • Saad A, Herrmann SMS, Eirin A, Ferguson CM, Glockner JF, Bjarnason H, McKusick MA, Misra S, Lerman LO, Textor SC. 2017. Phase 2a clinical trial of mitochondrial protection (Elamipretide) during stent revascularization in patients with atherosclerotic renal artery stenosis. Circ Cardiovasc Interv. 10(9):e005487.
  • Sabbah HN, Gupta RC, Kohli S, Wang M, Hachem S, Zhang K. 2016. Chronic therapy with Elamipretide (MTP-131), a novel mitochondria-targeting peptide, improves left ventricular and mitochondrial function in dogs with advanced heart failure. Circ Heart Fail. 9(2):e002206.
  • Safford SE, Oberley TD, Urano M, St Clair DK. 1994. Suppression of fibrosarcoma metastasis by elevated expression of manganese superoxide dismutase. Cancer Res. 54(16):4261–4265.
  • Salminen LE, Schofield PR, Pierce KD, Bruce SE, Griffin MG, Tate DF, Cabeen RP, Laidlaw DH, Conturo TE, Bolzenius JD, et al. 2017. Vulnerability of white matter tracts and cognition to the SOD2 polymorphism: a preliminary study of antioxidant defense genes in brain aging. Behav Brain Res. 329:111–119.
  • Sandbach JM, Coscun PE, Grossniklaus HE, Kokoszka JE, Newman NJ, Wallace DC. 2001. Ocular pathology in mitochondrial superoxide dismutase (Sod2)-deficient mice. Invest Ophthalmol Vis Sci. 42(10):2173–2178.
  • Sanjuan-Pla A, Cervera AM, Apostolova N, Garcia-Bou R, Victor VM, Murphy MP, McCreath KJ. 2005. A targeted antioxidant reveals the importance of mitochondrial reactive oxygen species in the hypoxic signaling of HIF-1alpha. FEBS Lett. 579(12):2669–2674.
  • Santini E, Turner KL, Ramaraj AB, Murphy MP, Klann E, Kaphzan H. 2015. Mitochondrial superoxide contributes to hippocampal synaptic dysfunction and memory deficits in Angelman syndrome model mice. J Neurosci. 35(49):16213–16220.
  • Sarsour EH, Kalen AL, Xiao Z, Veenstra TD, Chaudhuri L, Venkataraman S, Reigan P, Buettner GR, Goswami PC. 2012. Manganese superoxide dismutase regulates a metabolic switch during the mammalian cell cycle. Cancer Res. 72(15):3807–3816.
  • Saygi S, Erol I, Alehan F, Yalcin YY, Kubat G, Atac FB. 2015. Superoxide dismutase and catalase genotypes in pediatric migraine patients. J Child Neurol. 30(12):1586–1590.
  • Schaar CE, Dues DJ, Spielbauer KK, Machiela E, Cooper JF, Senchuk M, Hekimi S, Van Raamsdonk JM. 2015. Mitochondrial and cytoplasmic ROS have opposing effects on lifespan. PLoS Genet. 11(2):e1004972.
  • Schauen M, Spitkovsky D, Schubert J, Fischer JH, Hayashi J, Wiesner RJ. 2006. Respiratory chain deficiency slows down cell-cycle progression via reduced ROS generation and is associated with a reduction of p21CIP1/WAF1. J Cell Physiol. 209(1):103–112.
  • Scheurmann J, Treiber N, Weber C, Renkl AC, Frenzel D, Trenz-Buback F, Ruess A, Schulz G, Scharffetter-Kochanek K, Weiss JM. 2014. Mice with heterozygous deficiency of manganese superoxide dismutase (SOD2) have a skin immune system with features of “inflamm-aging. Arch Dermatol Res. 306(2):143–155.
  • Schriner SE, Linford NJ, Martin GM, Treuting P, Ogburn CE, Emond M, Coskun PE, Ladiges W, Wolf N, Van Remmen H, et al. 2005. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science. 308(5730):1909–1911.
  • Selvaratnam J, Robaire B. 2016. Overexpression of catalase in mice reduces age-related oxidative stress and maintains sperm production. Exp Gerontol. 84:12–20.
  • Sena LA, Chandel NS. 2012. Physiological roles of mitochondrial reactive oxygen species. Mol Cell. 48(2):158–167.
  • Sena LA, Li S, Jairaman A, Prakriya M, Ezponda T, Hildeman DA, Wang CR, Schumacker PT, Licht JD, Perlman H, et al. 2013. Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity. 38(2):225–236.
  • Shabalina IG, Vyssokikh MY, Gibanova N, Csikasz RI, Edgar D, Hallden-Waldemarson A, Rozhdestvenskaya Z, Bakeeva LE, Vays VB, Pustovidko AV, et al. 2017. Improved health-span and lifespan in mtDNA mutator mice treated with the mitochondrially targeted antioxidant SkQ1. Aging (Albany NY). 9(2):315–339.
  • Shadel GS, Horvath TL. 2015. Mitochondrial ROS signaling in organismal homeostasis. Cell. 163(3):560–569.
  • Shahzad K, Bock F, Dong W, Wang H, Kopf S, Kohli S, Al-Dabet MM, Ranjan S, Wolter J, Wacker C, et al. 2015. Nlrp3-inflammasome activation in non-myeloid-derived cells aggravates diabetic nephropathy. Kidney Int. 87(1):74–84.
  • Shen X, Zheng S, Metreveli NS, Epstein PN. 2006. Protection of cardiac mitochondria by overexpression of MnSOD reduces diabetic cardiomyopathy. Diabetes. 55(3):798–805.
  • Shetty S, Kumar R, Bharati S. 2019. Mito-TEMPO, a mitochondria-targeted antioxidant, prevents N-nitrosodiethylamine-induced hepatocarcinogenesis in mice. Free Radic Biol Med. 136:76–86.
  • Shi X, Bai Y, Zhang G, Liu Y, Xiao H, Liu X, Zhang W. 2018. Effects of over-expression of SOD2 in bone marrow-derived mesenchymal stem cells on traumatic brain injury. Cell Tissue Res. 372(1):67–75.
  • Shimoda-Matsubayashi S, Matsumine H, Kobayashi T, Nakagawa-Hattori Y, Shimizu Y, Mizuno Y. 1996. Structural dimorphism in the mitochondrial targeting sequence in the human manganese superoxide dismutase gene. A predictive evidence for conformational change to influence mitochondrial transport and a study of allelic association in Parkinson’s disease. Biochem Biophys Res Commun. 226(2):561–565.
  • Shimura T, Sasatani M, Kamiya K, Kawai H, Inaba Y, Kunugita N. 2016. Mitochondrial reactive oxygen species perturb AKT/cyclin D1 cell cycle signaling via oxidative inactivation of PP2A in lowdose irradiated human fibroblasts. Oncotarget. 7(3):3559–3570.
  • Siauciunaite R, Foulkes NS, Calabro V, Vallone D. 2019. Evolution shapes the gene expression response to oxidative stress. Int J Mol Sci. 20(12):3040.
  • Sidlauskaite E, Gibson JW, Megson IL, Whitfield PD, Tovmasyan A, Batinic-Haberle I, Murphy MP, Moult PR, Cobley JN. 2018. Mitochondrial ROS cause motor deficits induced by synaptic inactivity: Implications for synapse pruning. Redox Biol. 16:344–351.
  • Siegel MP, Kruse SE, Percival JM, Goh J, White CC, Hopkins HC, Kavanagh TJ, Szeto HH, Rabinovitch PS, Marcinek DJ. 2013. Mitochondrial-targeted peptide rapidly improves mitochondrial energetics and skeletal muscle performance in aged mice. Aging Cell. 12(5):763–771.
  • Sies H. 2014. Role of metabolic H2O2 generation: redox signaling and oxidative stress. J Biol Chem. 289(13):8735–8741.
  • Sies H, Jones DP. 2020. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat Rev Mol Cell Biol. 21(7):363–383.
  • Silachev DN, Plotnikov EY, Pevzner IB, Zorova LD, Balakireva AV, Gulyaev MV, Pirogov YA, Skulachev VP, Zorov DB. 2018. Neuroprotective effects of mitochondria-targeted plastoquinone in a rat model of neonatal hypoxic(-)ischemic brain injury. Molecules. 23(8):1871.
  • Silachev DN, Plotnikov EY, Zorova LD, Pevzner IB, Sumbatyan NV, Korshunova GA, Gulyaev MV, Pirogov YA, Skulachev VP, Zorov DB. 2015. Neuroprotective effects of mitochondria-targeted plastoquinone and thymoquinone in a rat model of brain ischemia/reperfusion injury. Molecules. 20(8):14487–14503.
  • Silva JP, Shabalina IG, Dufour E, Petrovic N, Backlund EC, Hultenby K, Wibom R, Nedergaard J, Cannon B, Larsson NG. 2005. SOD2 overexpression: enhanced mitochondrial tolerance but absence of effect on UCP activity. Embo J. 24(23):4061–4070.
  • Sims CR, MacMillan-Crow LA, Mayeux PR. 2014. Targeting mitochondrial oxidants may facilitate recovery of renal function during infant sepsis. Clin Pharmacol Ther. 96(6):662–664.
  • Skulachev VP. 2013. Cationic antioxidants as a powerful tool against mitochondrial oxidative stress. Biochem Biophys Res Commun. 441(2):275–279.
  • Skulachev VP, Anisimov VN, Antonenko YN, Bakeeva LE, Chernyak BV, Erichev VP, Filenko OF, Kalinina NI, Kapelko VI, Kolosova NG, et al. 2009. An attempt to prevent senescence: a mitochondrial approach. Biochim Biophys Acta. 1787(5):437–461.
  • Slanger TE, Chang-Claude J, Wang-Gohrke S. 2006. Manganese superoxide dismutase Ala-9Val polymorphism, environmental modifiers, and risk of breast cancer in a German population. Cancer Causes Control. 17(8):1025–1031.
  • Smith RA, Murphy MP. 2010. Animal and human studies with the mitochondria-targeted antioxidant MitoQ. Ann NY Acad Sci. 1201:96–103.
  • Smuder AJ, Sollanek KJ, Nelson WB, Min K, Talbert EE, Kavazis AN, Hudson MB, Sandri M, Szeto HH, Powers SK. 2018. Crosstalk between autophagy and oxidative stress regulates proteolysis in the diaphragm during mechanical ventilation. Free Radic Biol Med. 115:179–190.
  • Snow BJ, Rolfe FL, Lockhart MM, Frampton CM, O’Sullivan JD, Fung V, Smith RA, Murphy MP, Taylor KM. 2010. A double-blind, placebo-controlled study to assess the mitochondria-targeted antioxidant MitoQ as a disease-modifying therapy in Parkinson’s disease. Mov Disord. 25(11):1670–1674.
  • Soberanes S, Misharin AV, Jairaman A, Morales-Nebreda L, McQuattie-Pimentel AC, Cho T, Hamanaka RB, Meliton AY, Reyfman PA, Walter JM, et al. 2019. Metformin targets mitochondrial electron transport to reduce air-pollution-induced thrombosis. Cell Metab. 29(2):335–347. e335.
  • Soerensen M, Christensen K, Stevnsner T, Christiansen L. 2009. The Mn-superoxide dismutase single nucleotide polymorphism rs4880 and the glutathione peroxidase 1 single nucleotide polymorphism rs1050450 are associated with aging and longevity in the oldest old. Mech Ageing Dev. 130(5):308–314.
  • Solesio ME, Prime TA, Logan A, Murphy MP, Del Mar Arroyo-Jimenez M, Jordan J, Galindo MF. 2013. The mitochondria-targeted anti-oxidant MitoQ reduces aspects of mitochondrial fission in the 6-OHDA cell model of Parkinson’s disease. Biochim Biophys Acta. 1832(1):174–182.
  • Someya S, Xu J, Kondo K, Ding D, Salvi RJ, Yamasoba T, Rabinovitch PS, Weindruch R, Leeuwenburgh C, Tanokura M, et al. 2009. Age-related hearing loss in C57BL/6J mice is mediated by Bak-dependent mitochondrial apoptosis. Proc Natl Acad Sci USA. 106(46):19432–19437.
  • Sommer N, Huttemann M, Pak O, Scheibe S, Knoepp F, Sinkler C, Malczyk M, Gierhardt M, Esfandiary A, Kraut S, et al. 2017. Mitochondrial complex IV subunit 4 isoform 2 is essential for acute pulmonary oxygen sensing. Circ Res. 121(4):424–438.
  • Song MJ, Davidovich N, Lawrence GG, Margulies SS. 2016. Superoxide mediates tight junction complex dissociation in cyclically stretched lung slices. J Biomech. 49(8):1330–1335.
  • Song JQ, Jiang LY, Fu CP, Wu X, Liu ZL, Xie L, Wu XD, Hao SY, Li SQ. 2020. Heterozygous SOD2 deletion deteriorated chronic intermittent hypoxia-induced lung inflammation and vascular remodeling through mtROS-NLRP3 signaling pathway. Acta Pharmacol Sin. 41(9):1197–1207.
  • Song Y, Li S, Geng W, Luo R, Liu W, Tu J, Wang K, Kang L, Yin H, Wu X, et al. 2018. Sirtuin 3-dependent mitochondrial redox homeostasis protects against AGEs-induced intervertebral disc degeneration. Redox Biol. 19:339–353.
  • Souiden Y, Mallouli H, Meskhi S, Chaabouni Y, Rebai A, Cheour F, Mahdouani K. 2016. MnSOD and GPx1 polymorphism relationship with coronary heart disease risk and severity. Biol Res. 49:22.
  • Soule BP, Hyodo F, Matsumoto K, Simone NL, Cook JA, Krishna MC, Mitchell JB. 2007. The chemistry and biology of nitroxide compounds. Free Radic Biol Med. 42(11):1632–1650.
  • Sovari AA. 2011. Antioxidant therapy for atrial fibrillation: what is the next step? Cardiol Res Pract. 2011:429537.
  • Spisak K, Klimkowicz-Mrowiec A, Pera J, Dziedzic T, Aleksandra G, Slowik A. 2014. rs2070424 of the SOD1 gene is associated with risk of Alzheimer’s disease. Neurol Neurochir Pol. 48(5):342–345.
  • Spyrou G, Enmark E, Miranda-Vizuete A, Gustafsson J. 1997. Cloning and expression of a novel mammalian thioredoxin. J Biol Chem. 272(5):2936–2941.
  • Srivastava V, Buzas B, Momenan R, Oroszi G, Pulay AJ, Enoch MA, Hommer DW, Goldman D. 2010. Association of SOD2, a mitochondrial antioxidant enzyme, with gray matter volume shrinkage in alcoholics. Neuropsychopharmacology. 35(5):1120–1128.
  • Starkov AA. 2008. The role of mitochondria in reactive oxygen species metabolism and signaling. Ann NY Acad Sci. 1147:37–52.
  • Stclair D, Wan X, Kuroda M, Vichitbandha S, Tsuchida E, Urano M. 1997. Suppression of tumor metastasis by manganese superoxide dismutase is associated with reduced tumorigenicity and elevated fibronectin. Oncol Rep. 4(4):753–757.
  • Stealth Biotherapeutics 2019. [accessed 2020 Oct 13]. https://www.prnewswire.com/news-releases/stealth-biotherapeutics-provides-update-on-phase-3-trial-of-elamipretide-in-primary-mitochondrial-myopathy-300978082.html
  • Stefanova NA, Muraleva NA, Maksimova KY, Rudnitskaya EA, Kiseleva E, Telegina DV, Kolosova NG. 2016. An antioxidant specifically targeting mitochondria delays progression of Alzheimer’s disease-like pathology. Aging (Albany NY). 8(11):2713–2733.
  • Stelmashook EV, Stavrovskaya AV, Yamshchikova NG, Ol’shanskii AS, Kapay NA, Popova OV, Khaspekov LG, Skrebitsky VG, Isaev NK. 2015. Mitochondria-targeted plastoquinone antioxidant SkQR1 has positive effect on memory of rats. Biochemistry Mosc. 80(5):592–595.
  • Stepanova A, Konrad C, Manfredi G, Springett R, Ten V, Galkin A. 2019. The dependence of brain mitochondria reactive oxygen species production on oxygen level is linear, except when inhibited by antimycin A. J Neurochem. 148(6):731–745.
  • Stessman J, Maaravi Y, Hammerman-Rozenberg R, Cohen A, Nemanov L, Gritsenko I, Gruberman N, Ebstein RP. 2005. Candidate genes associated with ageing and life expectancy in the Jerusalem longitudinal study. Mech Ageing Dev. 126(2):333–339.
  • Stewart SF, Leathart JB, Chen Y, Daly AK, Rolla R, Vay D, Mottaran E, Vidali M, Albano E, Day CP. 2002. Valine-alanine manganese superoxide dismutase polymorphism is not associated with alcohol-induced oxidative stress or liver fibrosis. Hepatology. 36(6):1355–1360.
  • Stoehlmacher J, Ingles SA, Park DJ, Zhang W, Lenz HJ. 2002. The -9Ala/-9Val polymorphism in the mitochondrial targeting sequence of the manganese superoxide dismutase gene (MnSOD) is associated with age among Hispanics with colorectal carcinoma. Oncol Rep. 9(2):235–238.
  • St-Pierre J, Buckingham JA, Roebuck SJ, Brand MD. 2002. Topology of superoxide production from different sites in the mitochondrial electron transport chain. J Biol Chem. 277(47):44784–44790.
  • Strassburger M, Bloch W, Sulyok S, Schuller J, Keist AF, Schmidt A, Wenk J, Peters T, Wlaschek M, Lenart J, et al. 2005. Heterozygous deficiency of manganese superoxide dismutase results in severe lipid peroxidation and spontaneous apoptosis in murine myocardium in vivo. Free Radic Biol Med. 38(11):1458–1470.
  • Strauss E, Tomczak J, Staniszewski R, Oszkinis G. 2018. Associations and interactions between variants in selenoprotein genes, selenoprotein levels and the development of abdominal aortic aneurysm, peripheral arterial disease, and heart failure. PLoS One. 13(9):e0203350.
  • Stucki DM, Ruegsegger C, Steiner S, Radecke J, Murphy MP, Zuber B, Saxena S. 2016. Mitochondrial impairments contribute to Spinocerebellar ataxia type 1 progression and can be ameliorated by the mitochondria-targeted antioxidant MitoQ. Free Radic Biol Med. 97:427–440.
  • Sui Y, Fan Q, Wang B, Wang J, Chang Q. 2018. Ice-free cryopreservation of heart valve tissue: The effect of adding MitoQ to a VS83 formulation and its influence on mitochondrial dynamics. Cryobiology. 81:153–159.
  • Sukjamnong S, Chan YL, Zakarya R, Nguyen LT, Anwer AG, Zaky AA, Santiyanont R, Oliver BG, Goldys E, Pollock CA, et al. 2018. MitoQ supplementation prevent long-term impact of maternal smoking on renal development, oxidative stress and mitochondrial density in male mice offspring. Sci Rep. 8(1):6631.
  • Sullivan PG, Bruce-Keller AJ, Rabchevsky AG, Christakos S, Clair DK, Mattson MP, Scheff SW. 1999. Exacerbation of damage and altered NF-kappaB activation in mice lacking tumor necrosis factor receptors after traumatic brain injury. J Neurosci. 19(15):6248–6256.
  • Sun J, Folk D, Bradley TJ, Tower J. 2002. Induced overexpression of mitochondrial Mn-superoxide dismutase extends the life span of adult Drosophila melanogaster. Genetics. 161(2):661–672.
  • Sun L, Konig IR, Homann N. 2009. Manganese superoxide dismutase (MnSOD) polymorphism, alcohol, cigarette smoking and risk of oesophageal cancer. Alcohol Alcohol. 44(4):353–357.
  • Sun GG, Wang YD, Lu YF, Hu WN. 2013. Different association of manganese superoxide dismutase gene polymorphisms with risk of prostate, esophageal, and lung cancers: evidence from a meta-analysis of 20,025 subjects. Asian Pac J Cancer Prev. 14(3):1937–1943.
  • Supinski GS, Murphy MP, Callahan LA. 2009. MitoQ administration prevents endotoxin-induced cardiac dysfunction. Am J Physiol Regul Integr Comp Physiol. 297(4):R1095–R1102.
  • Suresh K, Servinsky L, Jiang H, Bigham Z, Yun X, Kliment C, Huetsch J, Damarla M, Shimoda LA. 2018. Reactive oxygen species induced Ca(2+) influx via TRPV4 and microvascular endothelial dysfunction in the SU5416/hypoxia model of pulmonary arterial hypertension. Am J Physiol Lung Cell Mol Physiol. 314(5):L893–L907.
  • Suresh K, Servinsky L, Jiang H, Bigham Z, Zaldumbide J, Huetsch JC, Kliment C, Acoba MG, Kirsch BJ, Claypool SM, et al. 2019. Regulation of mitochondrial fragmentation in microvascular endothelial cells isolated from the SU5416/hypoxia model of pulmonary arterial hypertension. Am J Physiol Lung Cell Mol Physiol. 317(5):L639–L652.
  • Sverdlov AL, Elezaby A, Qin F, Behring JB, Luptak I, Calamaras TD, Siwik DA, Miller EJ, Liesa M, Shirihai OS, et al. 2016. Mitochondrial reactive oxygen species mediate cardiac structural, functional, and mitochondrial consequences of diet-induced metabolic heart disease. J Am Heart Assoc. 5(1):e002555.
  • Sweetwyne MT, Pippin JW, Eng DG, Hudkins KL, Chiao YA, Campbell MD, Marcinek DJ, Alpers CE, Szeto HH, Rabinovitch PS, et al. 2017. The mitochondrial-targeted peptide, SS-31, improves glomerular architecture in mice of advanced age. Kidney Int. 91(5):1126–1145.
  • Szarka N, Pabbidi MR, Amrein K, Czeiter E, Berta G, Pohoczky K, Helyes Z, Ungvari Z, Koller A, Buki A, et al. 2018. Traumatic brain injury impairs myogenic constriction of cerebral arteries: role of mitochondria-derived H2O2 and TRPV4-dependent activation of BKca channels. J Neurotrauma. 35(7):930–939.
  • Szeto HH. 2006. Cell-permeable, mitochondrial-targeted, peptide antioxidants. Aaps J. 8(2):E277–E283.
  • Szeto HH. 2014. First-in-class cardiolipin-protective compound as a therapeutic agent to restore mitochondrial bioenergetics. Br J Pharmacol. 171(8):2029–2050.
  • Szeto HH, Birk AV. 2014. Serendipity and the discovery of novel compounds that restore mitochondrial plasticity. Clin Pharmacol Ther. 96(6):672–683.
  • Szeto HH, Liu S. 2018. Cardiolipin-targeted peptides rejuvenate mitochondrial function, remodel mitochondria, and promote tissue regeneration during aging. Arch Biochem Biophys. 660:137–148.
  • Szeto HH, Liu S, Soong Y, Alam N, Prusky GT, Seshan SV. 2016. Protection of mitochondria prevents high-fat diet-induced glomerulopathy and proximal tubular injury. Kidney Int. 90(5):997–1011.
  • Szeto HH, Liu S, Soong Y, Birk AV. 2015. Improving mitochondrial bioenergetics under ischemic conditions increases warm ischemia tolerance in the kidney. Am J Physiol Renal Physiol. 308(1):F11–21.
  • Szeto HH, Liu S, Soong Y, Seshan SV, Cohen-Gould L, Manichev V, Feldman LC, Gustafsson T. 2017. Mitochondria protection after acute ischemia prevents prolonged upregulation of IL-1beta and IL-18 and arrests CKD. J Am Soc Nephrol. 28(5):1437–1449.
  • Szeto HH, Liu S, Soong Y, Wu D, Darrah SF, Cheng FY, Zhao Z, Ganger M, Tow CY, Seshan SV. 2011. Mitochondria-targeted peptide accelerates ATP recovery and reduces ischemic kidney injury. J Am Soc Nephrol. 22(6):1041–1052.
  • Tabara LC, Poveda J, Martin-Cleary C, Selgas R, Ortiz A, Sanchez-Nino MD. 2014. Mitochondria-targeted therapies for acute kidney injury. Expert Rev Mol Med. 16:e13.
  • Takada Y, Hachiya M, Park SH, Osawa Y, Ozawa T, Akashi M. 2002. Role of reactive oxygen species in cells overexpressing manganese superoxide dismutase: mechanism for induction of radioresistance. Mol Cancer Res. 1(2):137–146.
  • Talbert EE, Smuder AJ, Min K, Kwon OS, Szeto HH, Powers SK. 2013. Immobilization-induced activation of key proteolytic systems in skeletal muscles is prevented by a mitochondria-targeted antioxidant. J Appl Physiol (1985). 115(4):529–538.
  • Tamer TM, Collins MN, Valachova K, Hassan MA, Omer AM, Mohy-Eldin MS, Svik K, Jurcik R, Ondruska L, Biro C, et al. 2018. MitoQ loaded chitosan-hyaluronan composite membranes for wound healing. Materials (Basel). 11(4):569.
  • Tamimi RM, Hankinson SE, Spiegelman D, Colditz GA, Hunter DJ. 2004. Manganese superoxide dismutase polymorphism, plasma antioxidants, cigarette smoking, and risk of breast cancer. Cancer Epidemiol Biomarkers Prev. 13(6):989–996.
  • Tarantini S, Valcarcel-Ares MN, Toth P, Yabluchanskiy A, Tucsek Z, Kiss T, Hertelendy P, Kinter M, Ballabh P, Sule Z, et al. 2019. Nicotinamide mononucleotide (NMN) supplementation rescues cerebromicrovascular endothelial function and neurovascular coupling responses and improves cognitive function in aged mice. Redox Biol. 24:101192.
  • Tarantini S, Valcarcel-Ares NM, Yabluchanskiy A, Fulop GA, Hertelendy P, Gautam T, Farkas E, Perz A, Rabinovitch PS, Sonntag WE, et al. 2018. Treatment with the mitochondrial-targeted antioxidant peptide SS-31 rescues neurovascular coupling responses and cerebrovascular endothelial function and improves cognition in aged mice. Aging Cell. 17(2):e12731.
  • Tate AD, Antonelli PJ, Hannabass KR, Dirain CO. 2017. Mitochondria-targeted antioxidant mitoquinone reduces cisplatin-induced ototoxicity in guinea pigs. Otolaryngol Head Neck Surg. 156(3):543–548.
  • Taufer M, Peres A, de Andrade VM, de Oliveira G, Sa G, do Canto ME, dos Santos AR, Bauer ME, da Cruz IB. 2005. Is the Val16Ala manganese superoxide dismutase polymorphism associated with the aging process? J Gerontol A Biol Sci Med Sci. 60(4):432–438.
  • Tefik T, Kucukgergin C, Sanli O, Oktar T, Seckin S, Ozsoy C. 2013. Manganese superoxide dismutase Ile58Thr, catalase C-262T and myeloperoxidase G-463A gene polymorphisms in patients with prostate cancer: relation to advanced and metastatic disease. BJU Int. 112(4):E406–414.
  • Teimoori B, Moradi-Shahrebabak M, Razavi M, Rezaei M, Harati-Sadegh M, Salimi S. 2019. The effect of GPx-1 rs1050450 and MnSOD rs4880 polymorphisms on PE susceptibility: a case- control study. Mol Biol Rep. 46(6):6099–6104.
  • Terry PD, Umbach DM, Taylor JA. 2005. No association between SOD2 or NQO1 genotypes and risk of bladder cancer. Cancer Epidemiol Biomarkers Prev. 14(3):753–754.
  • Thomas DA, Stauffer C, Zhao K, Yang H, Sharma VK, Szeto HH, Suthanthiran M. 2007. Mitochondrial targeting with antioxidant peptide SS-31 prevents mitochondrial depolarization, reduces islet cell apoptosis, increases islet cell yield, and improves posttransplantation function. J Am Soc Nephrol. 18(1):213–222.
  • Tian C, Fang S, Du X, Jia C. 2011. Association of the C47T polymorphism in SOD2 with diabetes mellitus and diabetic microvascular complications: a meta-analysis. Diabetologia. 54(4):803–811.
  • Tian C, Liu T, Fang S, Du X, Jia C. 2012. Association of C47T polymorphism in SOD2 gene with coronary artery disease: a case-control study and a meta-analysis. Mol Biol Rep. 39(5):5269–5276.
  • To EE, Erlich JR, Liong F, Luong R, Liong S, Esaq F, Oseghale O, Anthony D, McQualter J, Bozinovski S, et al. 2020. Mitochondrial reactive oxygen species contribute to pathological inflammation during influenza A virus infection in mice. Antioxid Redox Signal. 32(13):929–942.
  • Tomkins J, Banner SJ, McDermott CJ, Shaw PJ. 2001. Mutation screening of manganese superoxide dismutase in amyotrophic lateral sclerosis. Neuroreport. 12(11):2319–2322.
  • Townsend LK, Weber AJ, Barbeau PA, Holloway GP, Wright DC. 2020. Reactive oxygen species-dependent regulation of pyruvate dehydrogenase kinase-4 in white adipose tissue. Am J Physiol, Cell Physiol. 318(1):C137–C149.
  • Toyama S, Shimoyama N, Ishida Y, Koyasu T, Szeto HH, Shimoyama M. 2014. Characterization of acute and chronic neuropathies induced by oxaliplatin in mice and differential effects of a novel mitochondria-targeted antioxidant on the neuropathies. Anesthesiology. 120(2):459–473.
  • Toyama S, Shimoyama N, Szeto HH, Schiller PW, Shimoyama M. 2018. Protective effect of a mitochondria-targeted peptide against the development of chemotherapy-induced peripheral neuropathy in mice. ACS Chem Neurosci. 9(7):1566–1571.
  • Traba J, Kwarteng-Siaw M, Okoli TC, Li J, Huffstutler RD, Bray A, Waclawiw MA, Han K, Pelletier M, Sauve AA, et al. 2015. Fasting and refeeding differentially regulate NLRP3 inflammasome activation in human subjects. J Clin Invest. 125(12):4592–4600.
  • Treberg JR, Quinlan CL, Brand MD. 2011. Evidence for two sites of superoxide production by mitochondrial NADH-ubiquinone oxidoreductase (complex I). J Biol Chem. 286(31):27103–27110.
  • Treiber N, Maity P, Singh K, Kohn M, Keist AF, Ferchiu F, Sante L, Frese S, Bloch W, Kreppel F, et al. 2011. Accelerated aging phenotype in mice with conditional deficiency for mitochondrial superoxide dismutase in the connective tissue. Aging Cell. 10(2):239–254.
  • Treuting PM, Linford NJ, Knoblaugh SE, Emond MJ, Morton JF, Martin GM, Rabinovitch PS, Ladiges WC. 2008. Reduction of age-associated pathology in old mice by overexpression of catalase in mitochondria. J Gerontol A Biol Sci Med Sci. 63(8):813–822.
  • Trnka J, Blaikie FH, Smith RA, Murphy MP. 2008. A mitochondria-targeted nitroxide is reduced to its hydroxylamine by ubiquinol in mitochondria. Free Radic Biol Med. 44(7):1406–1419.
  • Tsybul’ko EA, Roshina NV, Rybina OY, Pasyukova EG. 2010. Mitochondria-targeted plastoquinone derivative SkQ1 increases early reproduction of Drosophila melanogaster at the cost of early survival. Biochemistry Moscow. 75(3):265–268.
  • Tuerdi A, Kinoshita M, Kamogashira T, Fujimoto C, Iwasaki S, Shimizu T, Yamasoba T. 2017. Manganese superoxide dismutase influences the extent of noise-induced hearing loss in mice. Neurosci Lett. 642:123–128.
  • Tuna A, Ozturk G, Gerceker TB, Karaca E, Onay H, Guvenc SM, Cogulu O. 2017. Superoxide dismutase 1 and 2 gene polymorphism in Turkish Vitiligo patients. Balkan J Med Genet. 20(2):67–74.
  • Turkseven S, Bolognesi M, Brocca A, Pesce P, Angeli P, Di Pascoli M. 2020. Mitochondria-targeted antioxidant mitoquinone attenuates liver inflammation and fibrosis in cirrhotic rats. Am J Physiol Gastrointest Liver Physiol. 318(2):G298–G304.
  • Turrens JF. 2003. Mitochondrial formation of reactive oxygen species. J Physiol (Lond). 552(Pt 2):335–344.
  • Ucer S, Iyer S, Kim H-N, Han L, Rutlen C, Allison K, Thostenson JD, de Cabo R, Jilka RL, O’Brien C, et al. 2017. The effects of aging and sex steroid deficiency on the murine skeleton are independent and mechanistically distinct. J Bone Miner Res. 32(3):560–574.
  • Ukai Y, Shiraishi M, Takeshita K, Fukui T, Kura K, Kimura K. 1988. Pharmacological studies on anethole trithione. Arzneimittelforschung. 38(10):1460–1465.
  • Ukai Y, Taniguchi N, Takeshita K, Kimura K, Enomoto H. 1984. Chronic anethole trithione treatment enhances the salivary secretion and increases the muscarinic acetylcholine receptors in the rat submaxillary gland. Arch Int Pharmacodyn Ther. 271(2):206–212.
  • Ukai Y, Taniguchi N, Takeshita K, Ogasawara T, Kimura K. 1988. Enhancement of salivary secretion by chronic anethole trithione treatment. Arch Int Pharmacodyn Ther. 294:248–258.
  • Ukai Y, Taniguchi N, Yamazaki A, Kimura K. 1989. Enhancement of phosphatidylinositol turnover and cyclic nucleotide accumulation by chronic anethole trithione treatment in rat submaxillary glands. J Pharm Pharmacol. 41(4):247–252.
  • Umanskaya A, Santulli G, Xie W, Andersson DC, Reiken SR, Marks AR. 2014. Genetically enhancing mitochondrial antioxidant activity improves muscle function in aging. Proc Natl Acad Sci USA. 111(42):15250–15255.
  • Underwood BR, Imarisio S, Fleming A, Rose C, Krishna G, Heard P, Quick M, Korolchuk VI, Renna M, Sarkar S, et al. 2010. Antioxidants can inhibit basal autophagy and enhance neurodegeneration in models of polyglutamine disease. Hum Mol Genet. 19(17):3413–3429.
  • Usui S, Komeima K, Lee SY, Jo YJ, Ueno S, Rogers BS, Wu Z, Shen J, Lu L, Oveson BC, et al. 2009. Increased expression of catalase and superoxide dismutase 2 reduces cone cell death in retinitis pigmentosa. Mol Ther. 17(5):778–786.
  • Vaka VR, McMaster KM, Cunningham MW, Jr, Ibrahim T, Hazlewood R, Usry N, Cornelius DC, Amaral LM, LaMarca B. 2018. Role of mitochondrial dysfunction and reactive oxygen species in mediating hypertension in the reduced uterine perfusion pressure rat model of preeclampsia. Hypertension. 72(3):703–711.
  • Valcarcel-Ares MN, Riveiro-Naveira RR, Vaamonde-Garcia C, Loureiro J, Hermida-Carballo L, Blanco FJ, Lopez-Armada MJ. 2014. Mitochondrial dysfunction promotes and aggravates the inflammatory response in normal human synoviocytes. Rheumatology (Oxford). 53(7):1332–1343.
  • Valenti L, Conte D, Piperno A, Dongiovanni P, Fracanzani AL, Fraquelli M, Vergani A, Gianni C, Carmagnola L, Fargion S. 2004. The mitochondrial superoxide dismutase A16V polymorphism in the cardiomyopathy associated with hereditary haemochromatosis. J Med Genet. 41(12):946–950.
  • van ‘t Erve TJ, Kadiiska MB, London SJ, Mason RP. 2017. Classifying oxidative stress by F2-isoprostane levels across human diseases: A meta-analysis. Redox Biol. 12:582–599.
  • van Golen RF, Reiniers MJ, Marsman G, Alles LK, van Rooyen DM, Petri B, Van der Mark VA, van Beek AA, Meijer B, Maas MA, et al. 2019. The damage-associated molecular pattern HMGB1 is released early after clinical hepatic ischemia/reperfusion. Biochim Biophys Acta Mol Basis Dis. 1865(6):1192–1200.
  • Van Raamsdonk JM, Hekimi S. 2009. Deletion of the mitochondrial superoxide dismutase sod-2 extends lifespan in Caenorhabditis elegans. PLoS Genet. 5(2):e1000361.
  • Van Remmen H, Ikeno Y, Hamilton M, Pahlavani M, Wolf N, Thorpe SR, Alderson NL, Baynes JW, Epstein CJ, Huang TT, et al. 2003. Life-long reduction in MnSOD activity results in increased DNA damage and higher incidence of cancer but does not accelerate aging. Physiol Genomics. 16(1):29–37.
  • Van Remmen H, Williams MD, Guo Z, Estlack L, Yang H, Carlson EJ, Epstein CJ, Huang TT, Richardson A. 2001. Knockout mice heterozygous for Sod2 show alterations in cardiac mitochondrial function and apoptosis. Am J Physiol Heart Circ Physiol. 281(3):H1422–H1432.
  • Vanita V. 2014. Association of RAGE (p.Gly82Ser) and MnSOD (p.Val16Ala) polymorphisms with diabetic retinopathy in T2DM patients from north India. Diabetes Res Clin Pract. 104(1):155–162.
  • Varga A, Budai MM, Milesz S, Bácsi A, Tőzsér J, Benkő S. 2013. Ragweed pollen extract intensifies lipopolysaccharide-induced priming of NLRP3 inflammasome in human macrophages. Immunology. 138(4):392–401.
  • Vats P, Sagar N, Singh TP, Banerjee M. 2015. Association of superoxide dismutases (SOD1 and SOD2) and glutathione peroxidase 1 (GPx1) gene polymorphisms with type 2 diabetes mellitus. Free Radic Res. 49(1):17–24.
  • Velando A, Noguera JC, da Silva A, Kim SY. 2019. Redox-regulation and life-history trade-offs: scavenging mitochondrial ROS improves growth in a wild bird. Sci Rep. 9(1):2203.
  • Velarde MC, Flynn JM, Day NU, Melov S, Campisi J. 2012. Mitochondrial oxidative stress caused by Sod2 deficiency promotes cellular senescence and aging phenotypes in the skin. Aging (Albany NY). 4(1):3–12.
  • Veldwijk MR, Herskind C, Sellner L, Radujkovic A, Laufs S, Fruehauf S, Zeller WJ, Wenz F. 2009. Normal-tissue radioprotection by overexpression of the copper-zinc and manganese superoxide dismutase genes. Strahlenther Onkol. 185(8):517–523.
  • Vendrov AE, Stevenson MD, Alahari S, Pan H, Wickline SA, Madamanchi NR, Runge MS. 2017. Attenuated superoxide dismutase 2 activity induces atherosclerotic plaque instability during aging in hyperlipidemic mice. J Am Heart Assoc. 6(11):e006775.
  • Vendrov AE, Vendrov KC, Smith A, Yuan J, Sumida A, Robidoux J, Runge MS, Madamanchi NR. 2015. NOX4 NADPH oxidase-dependent mitochondrial oxidative stress in aging-associated cardiovascular disease. Antioxid Redox Signal. 23(18):1389–1409.
  • Venkataraman S, Jiang X, Weydert C, Zhang Y, Zhang HJ, Goswami PC, Ritchie JM, Oberley LW, Buettner GR. 2005. Manganese superoxide dismutase overexpression inhibits the growth of androgen-independent prostate cancer cells. Oncogene. 24(1):77–89.
  • Venkataraman S, Wagner BA, Jiang X, Wang HP, Schafer FQ, Ritchie JM, Patrick BC, Oberley LW, Buettner GR. 2004. Overexpression of manganese superoxide dismutase promotes the survival of prostate cancer cells exposed to hyperthermia. Free Radic Res. 38(10):1119–1132.
  • Ventriglia M, Bocchio Chiavetto L, Scassellati C, Squitti R, Binetti G, Ghidoni R, Rossini PM, Gennarelli M. 2005. Lack of association between MnSOD gene polymorphism and sporadic Alzheimer’s disease. Aging Clin Exp Res. 17(6):445–448.
  • Ventriglia M, Scassellati C, Bonvicini C, Squitti R, Bevacqua MG, Foresti G, Tura GB, Gennarelli M. 2006. No association between Ala9Val functional polymorphism of MnSOD gene and schizophrenia in a representative Italian sample. Neurosci Lett. 410(3):208–211.
  • Vial G, Chauvin M-A, Bendridi N, Durand A, Meugnier E, Madec A-M, Bernoud-Hubac N, Pais de Barros J-P, Fontaine É, Acquaviva C, et al. 2015. Imeglimin normalizes glucose tolerance and insulin sensitivity and improves mitochondrial function in liver of a high-fat, high-sucrose diet mice model. Diabetes. 64(6):2254–2264.
  • Vicente-Gutierrez C, Bonora N, Bobo-Jimenez V, Jimenez-Blasco D, Lopez-Fabuel I, Fernandez E, Josephine C, Bonvento G, Enriquez JA, Almeida A, et al. 2019. Astrocytic mitochondrial ROS modulate brain metabolism and mouse behaviour. Nat Metab. 1(2):201–211.
  • Victorelli S, Lagnado A, Halim J, Moore W, Talbot D, Barrett K, Chapman J, Birch J, Ogrodnik M, Meves A, et al. 2019. Senescent human melanocytes drive skin ageing via paracrine telomere dysfunction. Embo J. 38(23):e101982.
  • Vilaseca M, Garcia-Caldero H, Lafoz E, Ruart M, Lopez-Sanjurjo CI, Murphy MP, Deulofeu R, Bosch J, Hernandez-Gea V, Gracia-Sancho J, et al. 2017. Mitochondria-targeted antioxidant mitoquinone deactivates human and rat hepatic stellate cells and reduces portal hypertension in cirrhotic rats. Liver Int. 37(7):1002–1012.
  • Vincent AM, Russell JW, Sullivan KA, Backus C, Hayes JM, McLean LL, Feldman EL. 2007. SOD2 protects neurons from injury in cell culture and animal models of diabetic neuropathy. Exp Neurol. 208(2):216–227.
  • Visscher H, Rassekh SR, Sandor GS, Caron HN, van Dalen EC, Kremer LC, van der Pal HJ, Rogers PC, Rieder MJ, Carleton BC, et al. 2015. Genetic variants in SLC22A17 and SLC22A7 are associated with anthracycline-induced cardiotoxicity in children. Pharmacogenomics. 16(10):1065–1076.
  • Vizioli MG, Liu T, Miller KN, Robertson NA, Gilroy K, Lagnado AB, Perez-Garcia A, Kiourtis C, Dasgupta N, Lei X, et al. 2020. Mitochondria-to-nucleus retrograde signaling drives formation of cytoplasmic chromatin and inflammation in senescence. Genes Dev. 34(5-6):428–445.
  • Vlachantoni D, Bramall AN, Murphy MP, Taylor RW, Shu X, Tulloch B, Van Veen T, Turnbull DM, McInnes RR, Wright AF. 2011. Evidence of severe mitochondrial oxidative stress and a protective effect of low oxygen in mouse models of inherited photoreceptor degeneration. Hum Mol Genet. 20(2):322–335.
  • von Montfort C, Matias N, Fernandez A, Fucho R, Conde de la Rosa L, Martinez-Chantar ML, Mato JM, Machida K, Tsukamoto H, Murphy MP, et al. 2012. Mitochondrial GSH determines the toxic or therapeutic potential of superoxide scavenging in steatohepatitis. J Hepatol. 57(4):852–859.
  • Wallace JL, Vaughan D, Dicay M, MacNaughton WK, de Nucci G. 2018. Hydrogen sulfide-releasing therapeutics: translation to the clinic. Antioxid Redox Signal. 28(16):1533–1540.
  • Wang Z, Cai F, Chen X, Luo M, Hu L, Lu Y. 2013. The role of mitochondria-derived reactive oxygen species in hyperthermia-induced platelet apoptosis. PLoS One. 8(9):e75044.
  • Wang DF, Cao B, Xu MY, Liu YQ, Yan LL, Liu R, Wang JY, Lu QB. 2015. Meta-analyses of manganese superoxide dismutase activity, gene Ala-9Val polymorphism, and the risk of schizophrenia. Medicine (Baltimore). 94(36):e1507.
  • Wang V, Chen SY, Chuang TC, Shan DE, Soong BW, Kao MC. 2010. Val-9Ala and Ile + 58Thr polymorphism of MnSOD in Parkinson’s disease. Clin Biochem. 43(12):979–982.
  • Wang W, Jin Y, Zeng N, Ruan Q, Qian F. 2017. SOD2 facilitates the antiviral innate immune response by scavenging reactive oxygen species. Viral Immunol. 30(8):582–589.
  • Wang A, Keita AV, Phan V, McKay CM, Schoultz I, Lee J, Murphy MP, Fernando M, Ronaghan N, Balce D, et al. 2014. Targeting mitochondria-derived reactive oxygen species to reduce epithelial barrier dysfunction and colitis. Am J Pathol. 184(9):2516–2527.
  • Wang M, Kirk JS, Venkataraman S, Domann FE, Zhang HJ, Schafer FQ, Flanagan SW, Weydert CJ, Spitz DR, Buettner GR, et al. 2005. Manganese superoxide dismutase suppresses hypoxic induction of hypoxia-inducible factor-1alpha and vascular endothelial growth factor. Oncogene. 24(55):8154–8166.
  • Wang J, Li J, Peng K, Fu ZY, Tang J, Yang MJ, Chen QC. 2017. Association of the C47T polymorphism in superoxide dismutase gene 2 with noise-induced hearing loss: a meta-analysis. Braz J Otorhinolaryngol. 83(1):80–87.
  • Wang H, Liu ZG, Peng J, Lin H, Zhong JX, Hu JY. 2009. The clinical therapic efficiency of anethol trithione on dry eye. Zhonghua Yan Ke Za Zhi. 45(6):492–497.
  • Wang G, McCain ML, Yang L, He A, Pasqualini FS, Agarwal A, Yuan H, Jiang D, Zhang D, Zangi L, et al. 2014. Modeling the mitochondrial cardiomyopathy of Barth syndrome with induced pluripotent stem cell and heart-on-chip technologies. Nat Med. 20(6):616–623.
  • Wang LI, Miller DP, Sai Y, Liu G, Su L, Wain JC, Lynch TJ, Christiani DC. 2001. Manganese superoxide dismutase alanine-to-valine polymorphism at codon 16 and lung cancer risk. J Natl Cancer Inst. 93(23):1818–1821.
  • Wang LI, Neuberg D, Christiani DC. 2004. Asbestos exposure, manganese superoxide dismutase (MnSOD) genotype, and lung cancer risk. J Occup Environ Med. 46(6):556–564.
  • Wang G, Shen T, Li P, Luo Z, Tan Y, He G, Zhang X, Yang J, Liu J, Wang Y, et al. 2019. The increase in IL-1β in the early stage of heatstroke might be caused by splenic lymphocyte pyroptosis induced by mtROS-mediated activation of the NLRP3 inflammasome. Front Immunol. 10:2862.
  • Wang JN, Shi N, Chen SY. 2012. Manganese superoxide dismutase inhibits neointima formation through attenuation of migration and proliferation of vascular smooth muscle cells. Free Radic Biol Med. 52(1):173–181.
  • Wang H, Sun X, Lin MS, Ferrario CM, Van Remmen H, Groban L. 2018. G protein-coupled estrogen receptor (GPER) deficiency induces cardiac remodeling through oxidative stress. Transl Res. 199:39–51.
  • Wang X, Tang D, Zou Y, Wu X, Chen Y, Li H, Chen S, Shi Y, Niu H. 2019. A mitochondrial-targeted peptide ameliorated podocyte apoptosis through a HOCl-alb-enhanced and mitochondria-dependent signalling pathway in diabetic rats and in vitro. J Enzyme Inhib Med Chem. 34(1):394–404.
  • Wang S, Wang F, Shi X, Dai J, Peng Y, Guo X, Wang X, Shen H, Hu Z. 2009. Association between manganese superoxide dismutase (MnSOD) Val-9Ala polymorphism and cancer risk – a meta-analysis. Eur J Cancer. 45(16):2874–2881.
  • Wang Y, Wang W, Wang N, Tall AR, Tabas I. 2017. Mitochondrial oxidative stress promotes atherosclerosis and neutrophil extracellular traps in aged mice. Arterioscler Thromb Vasc Biol. 37(8):e99–e107.
  • Wang P, Zhu Y, Xi S, Li S, Zhang Y. 2018. Association between MnSOD Val16Ala polymorphism and cancer risk: evidence from 33,098 cases and 37,831 controls. Dis Markers. 2018:3061974.
  • Wani WY, Gudup S, Sunkaria A, Bal A, Singh PP, Kandimalla RJ, Sharma DR, Gill KD. 2011. Protective efficacy of mitochondrial targeted antioxidant MitoQ against dichlorvos induced oxidative stress and cell death in rat brain. Neuropharmacology. 61(8):1193–1201.
  • Ward MS, Flemming NB, Gallo LA, Fotheringham AK, McCarthy DA, Zhuang A, Tang PH, Borg DJ, Shaw H, Harvie B, et al. 2017. Targeted mitochondrial therapy using MitoQ shows equivalent renoprotection to angiotensin converting enzyme inhibition but no combined synergy in diabetes. Sci Rep. 7(1):15190.
  • Wardman P. 2007. Fluorescent and luminescent probes for measurement of oxidative and nitrosative species in cells and tissues: progress, pitfalls, and prospects. Free Radic Biol Med. 43(7):995–1022.
  • Warnet JM, Christen MO, Thevenin M, Biard D, Jacqueson A, Claude JR. 1989a. Protective effect of anethol dithiolthione against acetaminophen hepatotoxicity in mice. Pharmacol Toxicol. 65(1):63–64.
  • Warnet JM, Christen MO, Thevenin M, Biard D, Jacqueson A, Claude JR. 1989b. Study of glutathione and glutathione related enzymes in acetaminophen-poisoned mice. Prevention by anethole trithione pretreatment. Arch Toxicol Suppl. 13:322–325.
  • Watson MA, Wong HS, Brand MD. 2019. Use of S1QELs and S3QELs to link mitochondrial sites of superoxide and hydrogen peroxide generation to physiological and pathological outcomes. Biochem Soc Trans. 47(5):1461–1469.
  • Watson MA, Khateeb S, Lopez-Dominguez J, Kapahi P, Brand MD. 2020. Mitochondrial superoxide from complex III drives diet-induced intestinal barrier dysfunction. Submitted for publication.
  • Wei Y, Troger A, Spahiu V, Perekhvatova N, Skulachev M, Petrov A, Chernyak B, Asbell P. 2019. The role of SKQ1 (Visomitin) in inflammation and wound healing of the ocular surface. Ophthalmol Ther. 8(1):63–73.
  • Weinberg F, Hamanaka R, Wheaton WW, Weinberg S, Joseph J, Lopez M, Kalyanaraman B, Mutlu GM, Budinger GR, Chandel NS. 2010. Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc Natl Acad Sci USA. 107(19):8788–8793.
  • Weisiger RA, Fridovich I. 1973. Mitochondrial superoxide dimutase. Site of synthesis and intramitochondrial localization. J Biol Chem. 248(13):4793–4796.
  • Wen JJ, Garg NJ. 2018. Manganese superoxide dismutase deficiency exacerbates the mitochondrial ROS production and oxidative damage in Chagas disease. PLoS Negl Trop Dis. 12(7):e0006687.
  • Wen Y, Liu Y, Tang T, Lv L, Liu H, Ma K, Liu B. 2016. NLRP3 inflammasome activation is involved in Ang II-induced kidney damage via mitochondrial dysfunction. Oncotarget. 7(34):54290–54302.
  • West AP, Brodsky IE, Rahner C, Woo DK, Erdjument-Bromage H, Tempst P, Walsh MC, Choi Y, Shadel GS, Ghosh S. 2011. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature. 472(7344):476–480.
  • Weydert C, Roling B, Liu J, Hinkhouse MM, Ritchie JM, Oberley LW, Cullen JJ. 2003. Suppression of the malignant phenotype in human pancreatic cancer cells by the overexpression of manganese superoxide dismutase. Mol Cancer Ther. 2(4):361–369.
  • Weydert CJ, Waugh TA, Ritchie JM, Iyer KS, Smith JL, Li L, Spitz DR, Oberley LW. 2006. Overexpression of manganese or copper-zinc superoxide dismutase inhibits breast cancer growth. Free Radic Biol Med. 41(2):226–237.
  • Wheatley-Price P, Asomaning K, Reid A, Zhai R, Su L, Zhou W, Zhu A, Ryan DP, Christiani DC, Liu G. 2008. Myeloperoxidase and superoxide dismutase polymorphisms are associated with an increased risk of developing pancreatic adenocarcinoma. Cancer. 112(5):1037–1042.
  • Wheeler MD, Katuna M, Smutney OM, Froh M, Dikalova A, Mason RP, Samulski RJ, Thurman RG. 2001. Comparison of the effect of adenoviral delivery of three superoxide dismutase genes against hepatic ischemia-reperfusion injury. Hum Gene Ther. 12(18):2167–2177.
  • Wheeler MD, Nakagami M, Bradford BU, Uesugi T, Mason RP, Connor HD, Dikalova A, Kadiiska M, Thurman RG. 2001. Overexpression of manganese superoxide dismutase prevents alcohol-induced liver injury in the rat. J Biol Chem. 276(39):36664–36672.
  • Wiegman CH, Michaeloudes C, Haji G, Narang P, Clarke CJ, Russell KE, Bao W, Pavlidis S, Barnes PJ, Kanerva J, et al. 2015. Oxidative stress-induced mitochondrial dysfunction drives inflammation and airway smooth muscle remodeling in patients with chronic obstructive pulmonary disease. J Allergy Clin Immunol. 136(3):769–780.
  • Wiener HW, Perry RT, Chen Z, Harrell LE, Go RC. 2007. A polymorphism in SOD2 is associated with development of Alzheimer’s disease. Genes Brain Behav. 6(8):770–775.
  • Wigner P, Czarny P, Synowiec E, Bijak M, Białek K, Talarowska M, Galecki P, Szemraj J, Sliwinski T. 2018. Variation of genes involved in oxidative and nitrosative stresses in depression. Eur Psychiatry. 48:38–48.
  • Williamson JR, Cooper RH. 1980. Regulation of the citric acid cycle in mammalian systems. FEBS Lett. 117(S1):K73–K85.
  • Wolf G, Aumann N, Michalska M, Bast A, Sonnemann J, Beck JF, Lendeckel U, Newsholme P, Walther R. 2010. Peroxiredoxin III protects pancreatic b cells from apoptosis. J Endocrinol. 207(2):163–175.
  • Wolf N, Penn P, Pendergrass W, Van Remmen H, Bartke A, Rabinovitch P, Martin GM. 2005. Age-related cataract progression in five mouse models for anti-oxidant protection or hormonal influence. Exp Eye Res. 81(3):276–285.
  • Wong GH. 1995. Protective roles of cytokines against radiation: induction of mitochondrial MnSOD. Biochim Biophys Acta. 1271(1):205–209.
  • Wong HS, Benoit B, Brand MD. 2018. Mitochondrial and cytosolic sources of hydrogen peroxide in resting C2C12 myoblasts. Free Radic Biol Med. 130:140–150.
  • Wong HS, Dighe PA, Mezera V, Monternier PA, Brand MD. 2017. Production of superoxide and hydrogen peroxide from specific mitochondrial sites under different bioenergetic conditions. J Biol Chem. 292(41):16804–16809.
  • Wong GH, Elwell JH, Oberley LW, Goeddel DV. 1989. Manganous superoxide dismutase is essential for cellular resistance to cytotoxicity of tumor necrosis factor. Cell. 58(5):923–931.
  • Wong HS, Monternier PA, Brand MD. 2019. S1QELs suppress mitochondrial superoxide/hydrogen peroxide production from site IQ without inhibiting reverse electron flow through Complex I. Free Radic Biol Med. 143:545–559.
  • Wong HS, Mezera V, Dighe P, Melov S, Gerencser AA, Sweis RF, Pliushchev M, Wang Z, Esbenshade T, McKibben B. 2020. Superoxide produced by mitochondrial site IQ inactivates cardiac succinate dehydrogenase and induces hepatic steatosis in Sod2 knockout mice. In preparation.
  • Wonsey DR, Zeller KI, Dang CV. 2002. The c-Myc target gene PRDX3 is required for mitochondrial homeostasis and neoplastic transformation. Proc Natl Acad Sci USA. 99(10):6649–6654.
  • Woo DK, Green PD, Santos JH, D’Souza AD, Walther Z, Martin WD, Christian BE, Chandel NS, Shadel GS. 2012. Mitochondrial genome instability and ROS enhance intestinal tumorigenesis in APC(Min/+) mice. Am J Pathol. 180(1):24–31.
  • Woodson K, Tangrea JA, Lehman TA, Modali R, Taylor KM, Snyder K, Taylor PR, Virtamo J, Albanes D. 2003. Manganese superoxide dismutase (MnSOD) polymorphism, alpha-tocopherol supplementation and prostate cancer risk in the alpha-tocopherol, beta-carotene cancer prevention study (Finland). Cancer Causes Control. 14(6):513–518.
  • Wu Y, Hao C, Liu X, Han G, Yin J, Zou Z, Zhou J, Xu C. 2020. MitoQ protects against liver injury induced by severe burn plus delayed resuscitation by suppressing the mtDNA-NLRP3 axis. Int Immunopharmacol. 80:106189.
  • Wu J, Hao S, Sun XR, Zhang H, Li H, Zhao H, Ji MH, Yang JJ, Li K. 2017. Elamipretide (SS-31) ameliorates isoflurane-induced long-term impairments of mitochondrial morphogenesis and cognition in developing rats. Front Cell Neurosci. 11:119.
  • Wu T, Huang Y, Gong Y, Xu Y, Lu J, Sheng H, Ni X. 2019. Treadmill exercise ameliorates depression-like behavior in the rats with prenatal dexamethasone exposure: the role of hippocampal mitochondria. Front Neurosci. 13:264.
  • Wu WB, Menon R, Xu YY, Zhao JR, Wang YL, Liu Y, Zhang HJ. 2016. Downregulation of peroxiredoxin-3 by hydrophobic bile acid induces mitochondrial dysfunction and cellular senescence in human trophoblasts. Sci Rep. 6:38946.
  • Wu J, Zhang M, Hao S, Jia M, Ji M, Qiu L, Sun X, Yang J, Li K. 2015. Mitochondria-targeted peptide reverses mitochondrial dysfunction and cognitive deficits in sepsis-associated encephalopathy. Mol Neurobiol. 52(1):783–791.
  • Xi Y, Feng D, Tao K, Wang R, Shi Y, Qin H, Murphy MP, Yang Q, Zhao G. 2018. MitoQ protects dopaminergic neurons in a 6-OHDA induced PD model by enhancing Mfn2-dependent mitochondrial fusion via activation of PGC-1alpha. Biochim Biophys Acta Mol Basis Dis. 1864(9):2859–2870.
  • Xiao Y, Meierhofer D. 2019. Are hydroethidine-based probes reliable for reactive oxygen species detection? Antioxid Redox Signal. 31(4):359–367.
  • Xiao L, Xu X, Zhang F, Wang M, Xu Y, Tang D, Wang J, Qin Y, Liu Y, Tang C, et al. 2017. The mitochondria-targeted antioxidant MitoQ ameliorated tubular injury mediated by mitophagy in diabetic kidney disease via Nrf2/PINK1. Redox Biol. 11:297–311.
  • Xie W, Santulli G, Reiken SR, Yuan Q, Osborne BW, Chen BX, Marks AR. 2015. Mitochondrial oxidative stress promotes atrial fibrillation. Sci Rep. 5:11427.
  • Xiong Y, Shie FS, Zhang J, Lee CP, Ho YS. 2005. Prevention of mitochondrial dysfunction in post-traumatic mouse brain by superoxide dismutase. J Neurochem. 95(3):732–744.
  • Xu L, Emery JF, Ouyang YB, Voloboueva LA, Giffard RG. 2010. Astrocyte targeted overexpression of Hsp72 or SOD2 reduces neuronal vulnerability to forebrain ischemia. Glia. 58(9):1042–1049.
  • Xu E, Liu J, Liu H, Wang X, Xiong H. 2018. Inflammasome activation by methamphetamine potentiates lipopolysaccharide stimulation of IL-1beta production in microglia. J Neuroimmune Pharmacol. 13(2):237–253.
  • Xu M, Wang L, Wang M, Wang H, Zhang H, Chen Y, Wang X, Gong J, Zhang JJ, Adcock IM, et al. 2019. Mitochondrial ROS and NLRP3 inflammasome in acute ozone-induced murine model of airway inflammation and bronchial hyperresponsiveness. Free Radic Res. 53(7):780–790.
  • Xu M, Xu M, Han L, Yuan C, Mei Y, Zhang H, Chen S, Sun K, Zhu B. 2017. Role for functional SOD2 polymorphism in pulmonary arterial hypertension in a Chinese population. Int J Environ Res Public Health. 14(3):266.
  • Xun Z, Rivera-Sanchez S, Ayala-Pena S, Lim J, Budworth H, Skoda EM, Robbins PD, Niedernhofer LJ, Wipf P, McMurray CT. 2012. Targeting of XJB-5-131 to mitochondria suppresses oxidative DNA damage and motor decline in a mouse model of Huntington’s disease. Cell Rep. 2(5):1137–1142.
  • Yahya MJ, Ismail PB, Nordin NB, Akim ABM, Binti Md Yusuf WS, Adam NLB, Zulkifli NF. 2019. CNDP1, NOS3, and MnSOD polymorphisms as risk factors for diabetic nephropathy among type 2 diabetic patients in Malaysia. J Nutr Metab. 2019:8736215.
  • Yan S, Brown SL, Kolozsvary A, Freytag SO, Lu M, Kim JH. 2008. Mitigation of radiation-induced skin injury by AAV2-mediated MnSOD gene therapy. J Gene Med. 10(9):1012–1018.
  • Yan L, Liu J, Wu S, Zhang S, Ji G, Gu A. 2014. Seminal superoxide dismutase activity and its relationship with semen quality and SOD gene polymorphism. J Assist Reprod Genet. 31(5):549–554.
  • Yan T, Oberley LW, Zhong W, St Clair DK. 1996. Manganese-containing superoxide dismutase overexpression causes phenotypic reversion in SV40-transformed human lung fibroblasts. Cancer Res. 56(12):2864–2871.
  • Yancey DM, Guichard JL, Ahmed MI, Zhou L, Murphy MP, Johnson MS, Benavides GA, Collawn J, Darley-Usmar V, Dell’Italia LJ. 2015. Cardiomyocyte mitochondrial oxidative stress and cytoskeletal breakdown in the heart with a primary volume overload. Am J Physiol Heart Circ Physiol. 308(6):H651–H663.
  • Yang LL, Huang MS, Huang CC, Wang TH, Lin MC, Wu CC, Wang CC, Lu SH, Yuan TY, Liao YH, et al. 2011. The association between adult asthma and superoxide dismutase and catalase gene activity. Int Arch Allergy Immunol. 156(4):373–380.
  • Yang HY, Jeong DK, Kim SH, Chung KJ, Cho EJ, Yang U, Lee SR, Lee TH. 2007. The role of peroxiredoxin III on late stage of proerythrocyte differentiation. Biochem Biophys Res Commun. 359(4):1030–1036.
  • Yang Y, Karakhanova S, Soltek S, Werner J, Philippov PP, Bazhin AV. 2012. In vivo immunoregulatory properties of the novel mitochondria-targeted antioxidant SkQ1. Mol Immunol. 52(1):19–29.
  • Yang SG, Park HJ, Kim JW, Jung JM, Kim MJ, Jegal HG, Kim IS, Kang MJ, Wee G, Yang HY, et al. 2018. Mito-TEMPO improves development competence by reducing superoxide in preimplantation porcine embryos. Sci Rep. 8(1):10130.
  • Yang Y, Yang L, Han Y, Wu Z, Chen P, Zhang H, Zhou J. 2017. Protective effects of hepatocyte-specific glycyrrhetic derivatives against carbon tetrachloride-induced liver damage in mice. Bioorg Chem. 72:42–50.
  • Yang L, Zhao K, Calingasan NY, Luo G, Szeto HH, Beal MF. 2009. Mitochondria targeted peptides protect against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine neurotoxicity. Antioxid Redox Signal. 11(9):2095–2104.
  • Yang S-G, Park H-J, Lee S-M, Kim J-W, Kim M-J, Kim I-S, Jegal H-G, Koo D-B. 2019. Reduction of mitochondrial derived superoxide by mito-TEMPO improves porcine oocyte maturation in vitro. J Anim Reprod Biotechnol. 34(1):10–19.
  • Yang Y, Xu P, Zhu F, Liao J, Wu Y, Hu M, Fu H, Qiao J, Lin L, Huang B. 2020. The potent antioxidant MitoQ protects against preeclampsia during late gestation but increases the risk of preeclampsia when administered in early pregnancy. Antioxid Redox Signal. doi:10.1089/ars.2019.7891. [Online Ahead of Print].
  • Yardeni T, Tanes CE, Bittinger K, Mattei LM, Schaefer PM, Singh LN, Wu GD, Murdock DG, Wallace DC. 2019. Host mitochondria influence gut microbiome diversity: a role for ROS. Sci Signal. 12(588):eaaw3159.
  • Ye J, Jiang Z, Chen X, Liu M, Li J, Liu N. 2017. The role of autophagy in pro-inflammatory responses of microglia activation via mitochondrial reactive oxygen species in vitro. J Neurochem. 142(2):215–230.
  • Yeh HL, Kuo LT, Sung FC, Yeh CC. 2018. Association between polymorphisms of antioxidant gene (MnSOD, CAT, and GPx1) and risk of coronary artery disease. Biomed Res Int. 2018:5086869.
  • Yen HC, Oberley TD, Vichitbandha S, Ho YS, St Clair DK. 1996. The protective role of manganese superoxide dismutase against adriamycin-induced acute cardiac toxicity in transgenic mice. J Clin Invest. 98(5):1253–1260.
  • Yin X, Manczak M, Reddy PH. 2016. Mitochondria-targeted molecules MitoQ and SS31 reduce mutant huntingtin-induced mitochondrial toxicity and synaptic damage in Huntington’s disease. Hum Mol Genet. 25(9):1739–1753.
  • Yoshikawa Y, Morita M, Hosomi H, Tsuneyama K, Fukami T, Nakajima M, Yokoi T. 2009. Knockdown of superoxide dismutase 2 enhances acetaminophen-induced hepatotoxicity in rat. Toxicology. 264(1–2):89–95.
  • Young ML, Franklin JL. 2019. The mitochondria-targeted antioxidant MitoQ inhibits memory loss, neuropathology, and extends lifespan in aged 3xTg-AD mice. Mol Cell Neurosci. 101:103409.
  • Yu WJ, Li NN, Tan EK, Cheng L, Zhang JH, Mao XY, Chang XL, Zhao DM, Liao Q, Peng R. 2014. No association of four candidate genetic variants in MnSOD and SYNIII with Parkinson’s disease in two Chinese populations. PLoS One. 9(2):e88050.
  • Zaha VG, Qi D, Su KN, Palmeri M, Lee HY, Hu X, Wu X, Shulman GI, Rabinovitch PS, Russell RR, et al. 2016. AMPK is critical for mitochondrial function during reperfusion after myocardial ischemia. J Mol Cell Cardiol. 91:104–113.
  • Zahreddine AM, Moustafa ME, Chamieh HA. 2016. Susceptibility of patients with manganese superoxide dismutase ala16val genetic polymorphism to type 2 Diabetes Mellitus and its complications in a sample of Lebanese population. J Genet Genome Res. 3(1):024.
  • Zai CC, Tiwari AK, Basile V, de Luca V, Muller DJ, Voineskos AN, Remington G, Meltzer HY, Lieberman JA, Potkin SG, et al. 2010. Oxidative stress in tardive dyskinesia: genetic association study and meta-analysis of NADPH quinine oxidoreductase 1 (NQO1) and Superoxide dismutase 2 (SOD2, MnSOD) genes. Prog Neuropsychopharmacol Biol Psychiatry. 34(1):50–56.
  • Zang QS, Sadek H, Maass DL, Martinez B, Ma L, Kilgore JA, Williams NS, Frantz DE, Wigginton JG, Nwariaku FE, et al. 2012. Specific inhibition of mitochondrial oxidative stress suppresses inflammation and improves cardiac function in a rat pneumonia-related sepsis model. Am J Physiol Heart Circ Physiol. 302(9):H1847–H1859.
  • Zejnilovic J, Akev N, Yilmaz H, Isbir T. 2009. Association between manganese superoxide dismutase polymorphism and risk of lung cancer. Cancer Genet Cytogenet. 189(1):1–4.
  • Zhan L, Li R, Sun Y, Dou M, Yang W, He S, Zhang Y. 2018. Effect of mito-TEMPO, a mitochondria-targeted antioxidant, in rats with neuropathic pain. Neuroreport. 29(15):1275–1281.
  • Zhang J, Bao X, Zhang M, Zhu Z, Zhou L, Chen Q, Zhang Q, Ma B. 2019. MitoQ ameliorates testis injury from oxidative attack by repairing mitochondria and promoting the Keap1-Nrf2 pathway. Toxicol Appl Pharmacol. 370:78–92.
  • Zhang XY, Chen DC, Xiu MH, Yang FD, Tan Y, Luo X, Zuo L, Kosten TA, Kosten TR. 2014. Cognitive function, plasma MnSOD activity, and MnSOD Ala-9Val polymorphism in patients with schizophrenia and normal controls. Schizophr Bull. 40(3):592–601.
  • Zhang X, Lu X, Li J, Xia Q, Gao J, Wu B. 2019. Mito-Tempo alleviates cryodamage by regulating intracellular oxidative metabolism in spermatozoa from asthenozoospermic patients. Cryobiology. 91:18–22.
  • Zhang L, Ma C, Zhang C, Ma M, Zhang F, Zhang L, Chen Y, Cao F, Li S, Zhu D. 2016. Reactive oxygen species effect PASMCs apoptosis via regulation of dynamin-related protein 1 in hypoxic pulmonary hypertension. Histochem Cell Biol. 146(1):71–84.
  • Zhang XY, Rao WW, Yu Q, Yu Y, Kou C, Tan YL, Chen DC, Zuo L, Luo X, Soares JC. 2016. Association of the manganese superoxide dismutase gene Ala-9Val polymorphism with age of smoking initiation in male schizophrenia smokers. Am J Med Genet B Neuropsychiatr Genet. 171B(2):243–249.
  • Zhang W, Tam J, Shinozaki K, Yin T, Lampe JW, Becker LB, Kim J. 2019. Increased survival time with SS-31 after prolonged cardiac arrest in rats. Heart Lung Circ. 28(3):505–508.
  • Zhang YG, Wang L, Kaifu T, Li J, Li X, Li L. 2016. Featured Article: accelerated decline of physical strength in peroxiredoxin-3 knockout mice. Exp Biol Med (Maywood). 241(13):1395–1400.
  • Zhang X, Wang Y, Wang H, Zhang M, Cai Z. 2017. The interaction between polymorphisms in MnSOD and prostate cancer risk: a meta-analysis and systematic review. Int J Clin Exp Med. 10:142–152.
  • Zhang J, Wang Q, Xu C, Lu Y, Hu H, Qin B, Wang Y, He D, Li C, Yu X, et al. 2017. MitoTEMPO prevents oxalate induced injury in NRK-52E cells via inhibiting mitochondrial dysfunction and modulating oxidative stress. Oxid Med Cell Longev. 2017:7528090.
  • Zhang T, Xu S, Wu P, Zhou K, Wu L, Xie Z, Xu W, Luo X, Li P, Ocak U, et al. 2019. Mitoquinone attenuates blood-brain barrier disruption through Nrf2/PHB2/OPA1 pathway after subarachnoid hemorrhage in rats. Exp Neurol. 317:1–9.
  • Zhang HJ, Yan T, Oberley TD, Oberley LW. 1999. Comparison of effects of two polymorphic variants of manganese superoxide dismutase on human breast MCF-7 cancer cell phenotype. Cancer Res. 59(24):6276–6283.
  • Zhang Y, Zhang C, Guan S, Zhou L, Li Q. 2017. Association between SOD2 rs6917589, rs2842980, rs5746136 and rs4880 polymorphisms and primary open angle glaucoma in a northern Chinese population. Int J Clin Exp Pathol. 10(3):3930–3938.
  • Zhang Z, Zhang X, Hou G, Sha W, Reynolds GP. 2002. The increased activity of plasma manganese superoxide dismutase in tardive dyskinesia is unrelated to the Ala-9Val polymorphism. J Psychiatr Res. 36(5):317–324.
  • Zhang Y, Zhang HM, Shi Y, Lustgarten M, Li Y, Qi W, Zhang BX, Van Remmen H. 2010. Loss of manganese superoxide dismutase leads to abnormal growth and signal transduction in mouse embryonic fibroblasts. Free Radic Biol Med. 49(8):1255–1262.
  • Zhang M, Zhao H, Cai J, Li H, Wu Q, Qiao T, Li K. 2017. Chronic administration of mitochondrion-targeted peptide SS-31 prevents atherosclerotic development in ApoE knockout mice fed Western diet. PLoS One. 12(9):e0185688.
  • Zhao WY, Han S, Zhang L, Zhu YH, Wang LM, Zeng L. 2013. Mitochondria-targeted antioxidant peptide SS31 prevents hypoxia/reoxygenation-induced apoptosis by down-regulating p66Shc in renal tubular epithelial cells. Cell Physiol Biochem. 32(3):591–600.
  • Zhao H, Li H, Hao S, Chen J, Wu J, Song C, Zhang M, Qiao T, Li K. 2017. Peptide SS-31 upregulates frataxin expression and improves the quality of mitochondria: implications in the treatment of Friedreich ataxia. Sci Rep. 7(1):9840.
  • Zhao K, Luo G, Giannelli S, Szeto HH. 2005. Mitochondria-targeted peptide prevents mitochondrial depolarization and apoptosis induced by tert-butyl hydroperoxide in neuronal cell lines. Biochem Pharmacol. 70(12):1796–1806.
  • Zhao W, Xu Z, Cao J, Fu Q, Wu Y, Zhang X, Long Y, Zhang X, Yang Y, Li Y, et al. 2019. Elamipretide (SS-31) improves mitochondrial dysfunction, synaptic and memory impairment induced by lipopolysaccharide in mice. J Neuroinflammation. 16(1):230.
  • Zhao K, Zhao GM, Wu D, Soong Y, Birk AV, Schiller PW, Szeto HH. 2004. Cell-permeable peptide antioxidants targeted to inner mitochondrial membrane inhibit mitochondrial swelling, oxidative cell death, and reperfusion injury. J Biol Chem. 279(33):34682–34690.
  • Zhong W, Oberley LW, Oberley TD, St Clair DK. 1997. Suppression of the malignant phenotype of human glioma cells by overexpression of manganese superoxide dismutase. Oncogene. 14(4):481–490.
  • Zhong W, Oberley LW, Oberley TD, Yan T, Domann FES, Clair DK. 1996. Inhibition of cell growth and sensitization to oxidative damage by overexpression of manganese superoxide dismutase in rat glioma cells. Cell Growth Differ. 7(9):1175–1186.
  • Zhong J, Xu C, Gabbay-Benziv R, Lin X, Yang P. 2016. Superoxide dismutase 2 overexpression alleviates maternal diabetes-induced neural tube defects, restores mitochondrial function and suppresses cellular stress in diabetic embryopathy. Free Radic Biol Med. 96:234–244.
  • Zhong W, Yan T, Webber MM, Oberley TD. 2004. Alteration of cellular phenotype and responses to oxidative stress by manganese superoxide dismutase and a superoxide dismutase mimic in RWPE-2 human prostate adenocarcinoma cells. Antioxid Redox Signaling. 6(3):513–522.
  • Zhou Y, Shuai P, Li X, Liu X, Wang J, Yang Y, Hao F, Lin H, Zhang D, Gong B. 2015. Association of SOD2 polymorphisms with primary open angle glaucoma in a Chinese population. Ophthalmic Genet. 36(1):43–49.
  • Zhou RH, Vendrov AE, Tchivilev I, Niu XL, Molnar KC, Rojas M, Carter JD, Tong H, Stouffer GA, Madamanchi NR, et al. 2012. Mitochondrial oxidative stress in aortic stiffening with age: the role of smooth muscle cell function. Arterioscler Thromb Vasc Biol. 32(3):745–755.
  • Zhou J, Wang H, Shen R, Fang J, Yang Y, Dai W, Zhu Y, Zhou M. 2018. Mitochondrial-targeted antioxidant MitoQ provides neuroprotection and reduces neuronal apoptosis in experimental traumatic brain injury possibly via the Nrf2-ARE pathway. Am J Transl Res. 10(6):1887–1899.
  • Zhou XL, Wei XJ, Li SP, Liu RN, Yu MX, Zhao Y. 2019. Interactions between cytosolic phospholipase A2 activation and mitochondrial reactive oxygen species production in the development of ventilator-induced diaphragm dysfunction. Oxid Med Cell Longev. 2019:2561929.
  • Zhu PR, Wu QY, Yu MM, Zhang MC, Ni MX, Liu SM, Jiang WJ, Zhang J, Li WW, Cao J, et al. 2017. Nucleotide polymorphism rs4880 of the SOD2 gene and the risk of male infertility. Zhonghua Nan Ke Xue. 23(2):137–141.
  • Zielonka J, Joseph J, Sikora A, Hardy M, Ouari O, Vasquez-Vivar J, Cheng G, Lopez M, Kalyanaraman B. 2017. Mitochondria-targeted triphenylphosphonium-based compounds: syntheses, mechanisms of action, and therapeutic and diagnostic applications. Chem Rev. 117(15):10043–10120.
  • Zorov DB, Juhaszova M, Sollott SJ. 2006. Mitochondrial ROS-induced ROS release: an update and review. Biochim Biophys Acta. 1757(5-6):509–517.
  • Zotova EV, Chistiakov DA, Savost’ianov KV, Bursa TR, Galeev IV, Strokov IA, Nosikov VV. 2003. Association of the SOD2 Ala(-9)Val and SOD3 Arg213Gly polymorphisms with diabetic polyneuropathy in patients with diabetes mellitus type 1. Mol Biol (Mosk). 37(3):404–408.