Home Progress in understanding the molecular oxygen paradox – function of mitochondrial reactive oxygen species in cell signaling
Article Publicly Available

Progress in understanding the molecular oxygen paradox – function of mitochondrial reactive oxygen species in cell signaling

  • Nidhi Kuksal , Julia Chalker and Ryan J. Mailloux EMAIL logo
Published/Copyright: July 4, 2017

Abstract

The molecular oxygen (O2) paradox was coined to describe its essential nature and toxicity. The latter characteristic of O2 is associated with the formation of reactive oxygen species (ROS), which can damage structures vital for cellular function. Mammals are equipped with antioxidant systems to fend off the potentially damaging effects of ROS. However, under certain circumstances antioxidant systems can become overwhelmed leading to oxidative stress and damage. Over the past few decades, it has become evident that ROS, specifically H2O2, are integral signaling molecules complicating the previous logos that oxyradicals were unfortunate by-products of oxygen metabolism that indiscriminately damage cell structures. To avoid its potential toxicity whilst taking advantage of its signaling properties, it is vital for mitochondria to control ROS production and degradation. H2O2 elimination pathways are well characterized in mitochondria. However, less is known about how H2O2 production is controlled. The present review examines the importance of mitochondrial H2O2 in controlling various cellular programs and emerging evidence for how production is regulated. Recently published studies showing how mitochondrial H2O2 can be used as a secondary messenger will be discussed in detail. This will be followed with a description of how mitochondria use S-glutathionylation to control H2O2 production.

Introduction

The production of reactive oxygen species (ROS) by aerobic systems is related to the unique chemical features of O2. Superoxide (O2˙) and hydrogen peroxide (H2O2) are the proximal ROS formed by mitochondria that can damage cell constituents at high levels. Fortunately, mitochondria (as well as the rest of the cell) are equipped with antioxidant systems that maintain O2˙ and H2O2 at low levels. Superoxide dismutase (SOD) is found in the intermembrane space (SOD1) and matrix (SOD2), where it rapidly converts two O2˙ to H2O2 which is then eliminated by a redundant series of enzyme systems (Halliwell, 1996). In most mitochondria, the glutathione (GSH) and thioredoxin-2 (TRX2) systems serve as the primary H2O2 removal systems, collectively referred to as the mitochondrial redox buffering system, which rely on the reductive power stored in NADPH for their reactivation (Deponte, 2013; Lu and Holmgren, 2014). Hepatocyte and heart mitochondria have also been reported to contain catalase, an additional line of defense that may be required to quench H2O2 at elevated concentrations (Radi et al., 1991; Drechsel and Patel, 2010).

The term oxidative stress was conceived when it was realized that antioxidant systems can become overburdened and/or disabled, leading to the accumulation of O2˙ and H2O2 culminating with cellular damage (Sies, 1985). The importance of antioxidant systems in preventing cellular damage is underscored by the consequences of genetically disabling different ROS eliminating enzymes in mitochondria. The genetic knockout of SOD2 or thioredoxin-2 (TRX2), which plays a vital role in PRX3 reactivation, is embryonically lethal (Hamilton et al., 2012). Loss of glutathione peroxidase-1 (GPX1) accelerates oxidative stress and damage and removing GPX4, which preferentially detoxifies lipid hydroperoxides, is also embryonically lethal (Ardanaz et al., 2010; Yoo et al., 2012). Exposure of mammals to O2 and its reactive intermediates is unavoidable. Molecular oxygen is the driving force behind oxidative phosphorylation (OXPHOS), the process by which mitochondria couple the combustion of nutrients and electron flow through a series of electron conducting multi-subunit enzyme complexes to the formation of ATP (reviewed in detail in Brand and Nicholls, 2011; Mailloux, 2015). It has recently been estimated that there can be up to 12 potential O2˙/H2O2 forming sites in mitochondria (reviewed in detail in Mailloux, 2015; Brand, 2016) and 31 ROS generating enzymes within the cell (Sies et al., 2017). In mitochondria, O2˙/H2O2 forming sites involved in metabolism include respiratory complexes I, II, and III and matrix-associated or membrane-bound dehydrogenases (Brand, 2016). Most of these O2˙/H2O2 forming sites are located in the matrix (Brand, 2016). However, two high capacity sites, sn-glycerol-3-phosphate dehydrogenase and complex III, emit O2˙/H2O2 directly into the intermembrane space (Brand, 2016). The O2˙/H2O2 forming capacity of the different sites varies substantially between different tissue types. In addition, several different factors influence how much ROS is formed from the different sites. These include membrane potential and mitochondrial bioenergetics, substrate type, the concentration of the carbon source being metabolized, the concentration of the electron donating site, accessibility of the electron donating site to O2, and the redox state of the electron donor (Murphy, 2009; Brand, 2016). Thus, it is essential for mitochondria to maintain a functional antioxidant system to ensure that ROS can be cleared and maintained at a low level.

Over the past few decades it has become evident that mitochondria do not have a simple relationship with ROS. Once considered unfortunate by-products of respiration, ROS, like H2O2, are now recognized as legitimate signaling molecules. In fact, H2O2 is sometimes referred to as a mitokine, a secondary messenger formed during normal mitochondrial metabolism that regulates cell function(s) (Yun and Finkel, 2014). It has also been hypothesized that O2˙, at low levels, fulfills a similar role – however, there is no convincing evidence to date that would suggest that it has a signaling function in mammals (James et al., 2012). Based on this, the definition for oxidative stress has since been revised to include terms such as oxidative eustress and oxidative distress (Sies et al., 2017). Oxidative eustress refers to a physiological oxidative challenge where H2O2 is utilized in adaptive or regulatory signaling (Sies et al., 2017). In this scenario, cellular ROS like H2O2 are utilized to oxidize redox switches on proteins to modulate genomic, transcriptional, and metabolic programs in response to changes in the exposome (Go et al., 2015). Thus, oxidative eustress, also known as the redox interface, allows for cellular adaptation in response to changes in the intracellular and extracellular environment (Jones, 2015). Oxidative distress on the other hand is associated with high ROS levels, oxidative damage, and the induction of apoptosis (Sies et al., 2017). Unlike oxidative eustress, protein oxidation is not reversible culminating with cellular dysfunction and death. The molecular oxygen paradox was coined to define the reciprocal relationship mammals have with O2; it is vital for aerobic respiration but the by-products associated with its use are disease-causing agents (Davies, 2016). However, it is now clear that ROS are not the price we pay for breathing but rather crucial signaling molecules that regulate mitochondrial function(s) in tandem with changes in physiology. This review examines the importance of mitochondrial H2O2, in regulating different cell programs. Recent studies that have delineated the importance of H2O2 in controlling cell stress, proliferation, and differentiation pathways will be discussed in detail. How mitochondria regulate O2˙/H2O2 production will also be examined in detail. In particular, this review will focus on S-glutathionylation reactions, a redox sensitive post-translational modifications that have been implicated in regulating mitochondrial O2˙/H2O2 production. We will discuss recent findings from our group and others showing that O2˙/H2O2 production by pyruvate dehydrogenase (PDH) and 2-oxoglutarate dehydrogenase (OGDH), two important sources of mitochondrial ROS, is regulated by protein S-glutathionylation. In addition, we discuss the function of OGDH as a GSH/GSSG sensor, adjusting its ROS output in response to changes in redox buffering capacity of the glutathione pool. The implication(s) of protein S-glutathionylation in modulating the secondary messenger function of H2O2 is also discussed in detail.

Mitokine function of hydrogen peroxide in mammalian cells

Interest in the physiological function of ROS can be traced back to the mid-1970s when it was discovered that oxygen is required to eliminate antigens in macrophages (Bernard et al., 1976). This was associated with the finding that tumor necrosis factor-α and interleukin-1, factors required for immune cell activation, stimulate O2˙ release in human fibroblasts (Meier et al., 1989); this increase in O2˙ production was associated with NADPH oxidase (NOX) activity (Meier et al., 1989). Collectively, these seminal findings demonstrated that ROS can fulfill important physiological functions, such as facilitating the degradation of pathogens by immune cells. The observation that O2˙ can fulfill a physiological function was considered to be a unique feature of phagocytes until it was found that moderately increased ROS levels can promote cell proliferation and growth (Nakamura et al., 1985). NOX was then discovered to be ubiquitously expressed leading to the thought that it may fulfill other physiological functions. Now, it is known that H2O2 formed following O2˙ production by NOX is required to activate cell signaling pathways and transcriptional programs (Rhee, 2006; Jones and Sies, 2015). In addition, different NOX isozymes are found in various cellular compartments including the cytosol, plasma membrane, and even in mitochondria (Graham et al., 2010). Thus, NOX isozymes are likely involved in the regulation of numerous cellular processes. Coupled with this, it was observed that ROS are required to trigger hypoxic signaling cascades and that mitochondria were required for this process (Chandel et al., 1998; Duranteau et al., 1998). It was eventually shown that the source of mitochondrial ROS was complex III and that release of ROS into the intermembrane space in low oxygen conditions induced hypoxic signaling (Bell et al., 2007). Which ROS is responsible for activating hypoxic signals is still a matter of debate. However, recent evidence would suggest that H2O2 is responsible for stabilizing HIF1α. Since then it has become evident that H2O2 emitted from mitochondria can potentially fulfill a number of cellular functions including stress signaling, apoptosis, T-cell activation, cell proliferation, adipocyte differentiation, steroidogenesis, and insulin sensitivity/resistance and release (Figure 1). The mitochondrial H2O2 mediated activation of the transcription factor NF-E2p45-related factor2 (Nrf2) is very well documented and has been shown to be integral for inducing the expression of antioxidant genes in response to electrophilic stress (Dinkova-Kostova and Abramov, 2015). In addition, the involvement of higher than normal mitochondrial H2O2 production in the activation of apoptotic cascades is also well characterized. In the following sections we will discuss new evidence showing that mitochondrial H2O2 fulfills other important cellular and physiological functions. Since the function of mitochondrial H2O2 in the activation of stress signaling cascades (hypoxia, Nrf2, and apoptosis) and cell proliferation is very well documented these pathways will not be discussed further.

Figure 1: H2O2 formed during nutrient metabolism can be utilized to activate a range of cell programs.Activation of various cell signaling cascades is dependent on the overall concentration of cellular H2O2. At low levels, H2O2 induces an adaptive signaling response associated with either the release of hormones, inhibition of steroidogenesis, modulation of circadian rhythms, immune cell activation, or induction of proliferation and differentiation. Slightly higher than normal levels triggers a stress signaling response. When H2O2 levels are too high, apoptosis pathways are activated through the induction of mitochondrial permeability transition.
Figure 1:

H2O2 formed during nutrient metabolism can be utilized to activate a range of cell programs.

Activation of various cell signaling cascades is dependent on the overall concentration of cellular H2O2. At low levels, H2O2 induces an adaptive signaling response associated with either the release of hormones, inhibition of steroidogenesis, modulation of circadian rhythms, immune cell activation, or induction of proliferation and differentiation. Slightly higher than normal levels triggers a stress signaling response. When H2O2 levels are too high, apoptosis pathways are activated through the induction of mitochondrial permeability transition.

T-cell activation

T-cells form a critical arm of the adaptive immune response. Antigen stimulation induces rapid T-cell division and differentiation into effector cells. T-cell activation is also accompanied by a dramatic shift in metabolism. Quiescent T-cells adopt a catabolic state where carbon is metabolized through the Krebs cycle, forming electron carriers that are oxidized to make ATP for basic housekeeping functions (Sena et al., 2013). T-cell activation results in a shift towards anabolic metabolism, where available carbon in the Krebs cycle is utilized to provide the constituents required for rapid cell division (Sena et al., 2013). ATP demands are met by an increase in glucose uptake and glycolysis. T-cell activation relies on the cross-linking of T-cell receptors (TCR), which induces a phosphorylation cascade that triggers the expression of factors involved in T-cell activation (Macian, 2005). Intriguingly, TCR activation is also associated with an increase in cell ROS production (Devadas et al., 2002). Through a set of ingenious experiments, Sena and coworkers found that this source of ROS was mitochondria, in particular complex III (Sena et al., 2013). Using mitochondria-targeted redox-sensitive GFP (mito-roGFP), it was found that T-cell activation results in increased mitochondrial H2O2 formation (Sena et al., 2013). Supplementation of cells with mitochondria-targeted antioxidants curtailed T-cell activation, which was also associated with attenuation of interleukin-2 production (Sena et al., 2013). In addition, T-cell-specific reduction of complex III curtailed mitochondrial ROS production and the induction of the adaptive immune response (Sena et al., 2013). It is important to note, however, that this study does not directly identify which mitochondrial ROS, O2˙ or H2O2, is required for T-cell activation. However, H2O2 is the most likely candidate because it has been found to modulate a number of transcription factors and proteins involved in various cellular processes through redox modification of cysteine residues (otherwise known as redox switches). ROS from complex III was required for NFAT activation and interleukin-2 production, two factors that are vital for T-cell activation and proliferation (Sena et al., 2013). In another study it was found that O2˙ production by complex I also stimulates NF-κB and AP1 signaling pathways, which are required to induce the expression of interleukins (Kaminski et al., 2010). Another intriguing aspect associated with T-cell activation is the upregulation of uncoupling protein-2 (UCP2), which plays a putative role in regulating mitochondrial O2˙/H2O2 production through the induction of proton return to the matrix. Antigen stimulation of T-cells is accompanied by an upregulation in UCP2 mRNA and protein expression (Chaudhuri et al., 2016). Quantitative expression analyses have revealed that UCP2 is mainly expressed in organs, cells, and tissues involved in the immune response (Rupprecht et al., 2012). Although speculative, it is possible that an increase in UCP2 expression after T-cell stimulation may be required to desensitize H2O2 signals once it has stimulated cell proliferation. This would allow mitochondria to control how much O2˙/H2O2 is formed, ensuring that the activated T-cell can take advantage its signaling properties whilst avoiding oxidative distress.

Adipocyte differentiation

White adipocytes are integral endocrine cells and organismal energy reservoirs. Adipocyte differentiation from mesenchymal stem cells (MSC) hinges on the activation of the transcription factors, C/EBPα and PPARγ. Stimulation of adipogenesis requires the activation of mammalian target of rapamycin (mTOR) complex-1 (mTORC1), which is mediated through the insulin and insulin-like growth factor stimulation of Akt signaling (Ricoult and Manning, 2013). Following its activation, mTORC1 elicits downstream effects that promote adipocyte differentiation (e.g. induction of C/EBPα and PPARγ). It has been found that mitochondrial respiration and ROS production also increase during adipogenesis (Wilson-Fritch et al., 2003; Imhoff and Hansen, 2010). In one particular study, it was found that stimulation of mTORC1 results in an increase in mitochondrial metabolism and biogenesis, which was also associated by an increase in O2˙/H2O2 production (Tormos et al., 2011). Differentiation of MSC correlated strongly with an increase in H2O2 levels, which could be inhibited with mitochondria-targeted antioxidants (Tormos et al., 2011). Moreover, MSC cultures treated with mitochondria-targeted antioxidants also diminished the expression of several transcriptional regulators that facilitate adipocyte differentiation (Tormos et al., 2011). Finally, supplementing MSC cultures exposed to mitochondria-targeted antioxidants with galactose oxidase/galactose, which produces H2O2, facilitated adipocyte differentiation and lipid accumulation. It was also shown that the elevation of cellular O2˙/H2O2 was an early event during adipogenesis pointing to its potential function in serving as a co-signal for differentiation (Tormos et al., 2011). It should be emphasized at this point that cellular H2O2 is well documented to play a role in cell differentiation (Weinberg et al., 2015). By knocking down Rieske iron sulfur protein (RISP), which prevents the accumulation of semiquinone and O2˙ formation, the authors were able to show that the source of ROS was complex III of the electron transport chain (Tormos et al., 2011). Thus, it would appear that mitochondrial H2O2 plays a role in reinforcing signaling programs in differentiating adipocytes. Hydrogen peroxide has also been found to induce the phosphorylation and activation of Akt and the autophosphorylation of insulin receptor, both of which play a part in the induction of adipocyte differentiation (Papaconstantinou, 2009; Sadidi et al., 2009). Thus, it is possible that mitochondrial H2O2 release serves as a positive feedback signal that prolongs the induction of adipocyte differentiation cascades. Interestingly, white adipocytes also express UCP2, which may be required to desensitize the ROS signal emitted by mitochondria (Anedda et al., 2008). A fairly recent study found that UCP2 displays variable expression in pluripotent stem cells and its upregulation limits H2O2 production, inhibiting cellular differentiation (Zhang et al., 2011). Thus, it is possible that UCP2 is either utilized to curtail adipogenesis or is expressed after differentiation to quench the ROS signal.

Steroidogenesis in the adrenal cortex and circadian rhythms

Peroxiredoxins are a family of H2O2 clearing enzymes that are well conserved from archaea to humans (Perkins et al., 2014). There are six known mammalian isozymes (PRX1-6), with PRX3 and PRX5 found in the matrix of mitochondria (Zhu et al., 2012). Aside from H2O2 degradation, PRX proteins have also been found to mediate cell signals. This can be achieved either by (1) serving as a redox relay where the disulfide formed on PRX is then transferred to a transcription factor (Sobotta et al., 2015) or (2) via the floodgate model where overoxidation of PRX prolongs its deactivation allowing for an accumulation of H2O2 for signaling (Jarvis et al., 2012). Although the former is intriguing, for our purposes here we will focus on mechanism number two since it involves mitochondria. The normal catalytic cycle of PRX involves the H2O2-mediated oxidation of a catalytic thiol, which forms a sulfenic acid (SOH) moiety that is quickly resolved by a second cysteine residue. This results in the formation of a disulfide bridge that is reduced in an NADPH- and TRX-dependent fashion. If H2O2 levels are high enough, SOH can be oxidized to a sulfinic acid (SO2H) (Rhee and Kil, 2017). Although SO2H groups can be reduced back to an active cysteine thiol by sulfiredoxin (SRX), the reaction is slow. Thus, the overoxidation of PRX can prolong its deactivation allowing H2O2 levels to accumulate eliciting a cell signaling response (Rhee and Kil, 2017). Once sulfiredoxin successfully reduces SO2H and reactivates PRX, cellular H2O2 can be cleared desensitizing the signal.

The reaction mechanism described above mostly applies to the cytosolic PRX isoforms, PRX1 and PRX2 (Rhee and Kil, 2017). However, recent work has found that this concept also applies to mitochondrial PRX isoform, PRX3. Kil and colleagues found that the over-oxidation of PRX3 is instrumental for the mitochondrial H2O2 mediated regulation of corticosterone production in the adrenal cortex (Kil et al., 2012). The production of glucocorticoids is activated by the adrenocorticotropic hormone (ACTH) response – ACTH secreted from the pituitary gland activates steroidogenesis in the adrenal cortex (Kil et al., 2012). Induction of steroidogenesis requires the production of cAMP, which stimulates the synthesis of corticosterone from cholesterol by activating cytochrome P450 (CYP) enzymes (Kil et al., 2012). Intriguingly, one of these CYP enzymes, called CYP11B1, is found in mitochondria where it catalyzes the last step of corticosterone synthesis. CYP enzymes are notorious sources of ROS. It was also found that oxidative stress can hinder steroidogenesis in the adrenal cortex by decreasing the availability of steroidogenic acute regulatory protein (StAR), which plays a vital role in the final step of corticosterone biosynthesis (Diemer et al., 2003). High ROS also activates p38 mitogen activated protein kinase, which inhibits StAR expression (Abidi et al., 2008). Bringing this full circle, Kil and coworkers found that, during steroidogenesis, the H2O2 formed by CYP11B1 in the matrix of mitochondria overoxidizes PRX3 (Kil et al., 2012). This allows H2O2 accumulation and diffusion into the cytosol where it activates p38 inhibiting StAR and corticosterone production (Kil et al., 2012). Therefore, H2O2 formed by CYP11B1 during steroidogenesis serves as a feedback inhibitor for corticosterone production. One potential issue associated with these findings is that PRX3, unlike PRX1, is highly resistant to overoxidation (Cox et al., 2009). A viable way to explain how PRX3 is deactivated is that the H2O2 levels become quite high during corticosterone production, allowing for PRX3 hyperoxidation (to our knowledge the absolute concentration of H2O2 in the adrenal cortex during steroidogenesis has not been quantified). This could indeed be the case since in the same study the authors found that ~20% of the PRX3 pool was in the sulfinic acid form whereas it was not detected in other tissues (Kil et al., 2012).

In the same study published by Kil and colleagues, it was also found that PRX3 oxidation displays a circadian pattern (Kil et al., 2012). In the adrenal cortex, the sulfinic acid form of PRX3 was at its lowest at 12 h; this coincided with the mitochondrial uptake of SRX and a decrease in phosphorylated p38 (Rhee and Kil, 2016). After 24 h, SRX levels decreased, which was matched by an increase in PRX3 hyperoxidation and p38 activation. Mitochondrial import of SRX relies on HSP90 and its subsequent removal is facilitated by Lon protease (Rhee and Kil, 2016). Comparable results were obtained with brown adipose and heart tissue, suggesting that various tissue types utilize PRX3 oxidation and H2O2 signaling as a common mechanism for the circadian rhythm-mediated regulation of cell physiology.

Glucose stimulated insulin release

Insulin is likely the most discussed hormone in the scientific and public sphere since lack of production is associated with the development of two of the most common metabolic disorders: type 1 and type 2 diabetes. This is simply because it serves as the linchpin for the regulation of overall body metabolism after a meal. It has been known for some time that insulin release from pancreatic β-cells is stimulated by a rise in blood glucose levels. Briefly, glucose is imported into the cytoplasm of β-cells after a meal and is metabolized through glycolysis and oxidative phosphorylation forming ATP, inducing insulin release (Jensen et al., 2008). Other secretagogues, like fatty acids and amino acids, have also been found to potentiate insulin release, mostly during the second phase insulin secretion (Andrikopoulos, 2016). Insulin secretion is also sensitive towards redox signals, in particular mitochondrial H2O2 (Leloup et al., 2009). Indeed, exposure of β-cells to high glucose (5–16.7 mm) induces a robust increase in cellular ROS and insulin release, a response that can be blunted by exogenous antioxidants (Leloup et al., 2009). Glucose-stimulated insulin release could also be simulated using respiratory chain inhibitors that increase cellular O2˙/H2O2 levels (Leloup et al., 2009). In addition, it was found that the O2˙ generator, paraquat, and exogenous H2O2 can also promote insulin release (Mailloux et al., 2012ab). The finding that H2O2 can serve as an insulin secretagogue is supported by the presence of UCP2 in pancreatic β-cells, which has been shown in several studies to attenuate glucose-stimulated insulin release and can account for up to ~70% of cellular respiration (Affourtit et al., 2011; Robson-Doucette et al., 2011). UCP2-mediated regulation of glucose-stimulated insulin release stems from its putative function in proton leaks. In this scenario, the UCP2-induced return of protons to the matrix diminishes the protonmotive force, lowering O2˙/H2O2 release by the respiratory chain. It was found later that during the initial uptake and metabolism glucose, UCP2 is maintained in an inactive state inducing a spike in H2O2 levels in cultured insulinoma cells and pancreatic islets, promoting insulin release (Mailloux et al., 2012ab). The increase in cellular H2O2 also feeds back on UCP2 activating proton leaks, desensitizing the insulin secretion signal (Affourtit et al., 2011). Inducible leaks through UCP2 and control over ROS production was later found to be associated with changes in the S-glutathionylation of the protein (discussed in detail below in the section ‘Regulation of mitochondrial hydrogen peroxide production’) (Mailloux et al., 2012ab). Indeed, the S-glutathionylation agents, diamide and BioGEE, were able to stimulate insulin secretion via the modification of UCP2 (Mailloux et al., 2012ab). Recent work has also shown that ROS and UCP2 also play a role in controlling glucagon release from pancreatic α-cells, indicating that the pancreas utilizes a common regulatory mechanism to control the secretion of two very important metabolic hormones into the blood stream (Allister et al., 2013). Although it is unknown if S-glutathionylation events are required to control insulin release in vivo, results collected so far indicate redox signals are required to modulate glucose-stimulated insulin release and the release of glucagon from α-cells during starvation.

Regulation of mitochondrial hydrogen peroxide production

Proton leaks

Mitochondrial O2˙/H2O2 levels can be regulated via degradation pathways or control over production (Mailloux et al., 2016a,b). O2˙ and H2O2 are degraded by a series of redundant antioxidant systems, which were briefly discussed in the introduction and reviewed extensively elsewhere (Murphy, 2012). Control over its production also represents an important means of modulating how much O2˙/H2O2 is available in a cell. Mechanisms for regulation include (1) proton leaks, (2) supercomplex assemblies, and (3) redox signals (e.g. protein S-glutathionylation) (Figure 2). Of these mechanisms, the best characterized is proton leaks. Proton leaks are defined as the return of protons to the mitochondrial matrix, by-passing ATP synthase (Jastroch et al., 2010). The function of proton leaks in regulating mitochondrial ROS production has been extensively reviewed (Harper et al., 2008; Divakaruni and Brand, 2011; Mailloux et al., 2011). Thus, we will only discuss this concept in brief whilst trying to highlight some of the debate surrounding how leaks can regulate ROS production. The main proteins involved in leaks are adenine nucleotide transporter (ANT) and uncoupling proteins (UCP) 1-5 (Divakaruni and Brand, 2011; Ramsden et al., 2012). A non-Ohmic relationship exists between the strength of the proton gradient and mitochondrial ROS production, where a small increase in protonmotive force can lead to the augmentation of mitochondrial O2˙/H2O2 production (Korshunov et al., 1997). Enhancing proton return to the matrix can have the opposite effect, diminishing ROS production by mitochondria. This has been shown with protonophores such as FCCP, which uncouple respiration diminishing ROS production. By decreasing protonic back pressure, successful electron transfer through the respiratory chain to O2 at complex IV increases, diminishing the amount of electrons available to form O2˙/H2O2 (Divakaruni and Brand, 2011). However, the use of proton leaks to govern how much ROS is actually formed by mitochondria has been a focal point for controversy for more than a decade. For example, it has been found that activation of leaks through UCP1 in brown fat mitochondria, either by oleate or cold acclimation, increases mitochondrial ROS production (Mailloux et al., 2012ab; Shabalina et al., 2014). It is also not entirely clear as to whether or not UCP2 (more ubiquitously expressed) and UCP3 (found exclusively in skeletal muscle) utilize leaks to prevent oxidative damage and control ROS production. For instance, it has been found that both putative uncouplers can eliminate lipid hydroperoxides (Lombardi et al., 2010). Some studies have found that UCP2 and UCP3 translocate calcium ions into the matrix of mitochondria whereas others have shown that UCP3 transports fatty acids (Schrauwen et al., 2003; Trenker et al., 2007). Moreover, recent work has suggested that mitochondrial O2˙/H2O2 production is independent of the strength of the protonmotive force as inhibition of the electron transport chain and decreasing the rate of respiration do not necessarily correlate with how much ROS is formed (Brand, 2016). In addition, increasing mitochondrial respiration rate and proton return can also augment ROS production (Mailloux et al., 2012ab). It is important to emphasize that numerous studies have shown that the uncoupling proteins, in particular UCP2 and UCP3, play vital roles in preventing oxidative stress and damage and fulfill some important physiological functions (e.g. optimization of fuel metabolism in muscle or satiety signaling and insulin release). However, whether these roles are related to the induction of proton leaks and control over mitochondrial O2˙/H2O2 release is still a matter for debate.

Figure 2: Regulation of mitochondrial H2O2 release.H2O2 formed during nutrient metabolism can be increased or decreased by proton leaks (i), supercomplex assemblies (ii), or redox signals (iii). (i) Induction of proton return by adenosine nucleotide translocator (ANT) or uncoupling proteins (UCP) lowers the strength of the protonmotive force across the mitochondrial inner membrane. This diminishes protonic back pressure on the electron transport chain, increasing electron flow through the chain and the successful reduction of O2 to H2O at complex IV, which decreases the number of electrons available for ROS production. (ii) Fully assembled respirasomes containing complex I, III, and IV with ubiquinone and cytochrome c diminish ROS production by increasing the efficiency of electron flow through the chain. Disassembly results in decreased efficiency in electron movement, over-reduction of electron donating sites, and increased ROS formation. (iii) The degree of protein S-glutathionylation in mitochondria is dictated by the concentration of H2O2 which alters the availability of GSH and GSSG. A decrease in the GSH/GSSG redox pair increases S-glutathionylation of ROS forming sites in mitochondria diminishing H2O2 formation.
Figure 2:

Regulation of mitochondrial H2O2 release.

H2O2 formed during nutrient metabolism can be increased or decreased by proton leaks (i), supercomplex assemblies (ii), or redox signals (iii). (i) Induction of proton return by adenosine nucleotide translocator (ANT) or uncoupling proteins (UCP) lowers the strength of the protonmotive force across the mitochondrial inner membrane. This diminishes protonic back pressure on the electron transport chain, increasing electron flow through the chain and the successful reduction of O2 to H2O at complex IV, which decreases the number of electrons available for ROS production. (ii) Fully assembled respirasomes containing complex I, III, and IV with ubiquinone and cytochrome c diminish ROS production by increasing the efficiency of electron flow through the chain. Disassembly results in decreased efficiency in electron movement, over-reduction of electron donating sites, and increased ROS formation. (iii) The degree of protein S-glutathionylation in mitochondria is dictated by the concentration of H2O2 which alters the availability of GSH and GSSG. A decrease in the GSH/GSSG redox pair increases S-glutathionylation of ROS forming sites in mitochondria diminishing H2O2 formation.

Supercomplexes

Other criteria for controlling mitochondrial O2˙/H2O2 production include (1) substrate supply, (2) the type of substrate being oxidized, (3) redox state of electron donating sites, (4) concentration of electron donating sites, and (5) access to oxygen. In this regard, the assembly of supercomplexes, also called respirasomes, represents one potential mechanism that may modulate O2˙/H2O2 production by controlling electron movement between different sites of ROS production (Enriquez, 2016). The respiratory supercomplex theory was originally developed to explain the co-migration of multi-respiratory complex units composed of complexes I, III, and IV during blue native polyacrylamide gel electrophoresis (Schagger and Pfeiffer, 2000). It was estimated that most of the complex I pool is involved in forming adducts with complex III, which form a functional “respirasome” with complex IV (Schagger and Pfeiffer, 2000). In 2008, Acin-Perez and colleagues found that blue native bands corresponding to these so-called respirasomes were enzymatically active and able to transfer electrons from NADH at complex I to complex IV resulting in a measurable rate of O2 consumption (Acin-Perez et al., 2008). Supercomplex assemblies have been found to vary widely in composition, leading to the development of the plasticity model for respirasome formation, which affects the overall respiratory efficiency of mitochondrial preparations (Moreno-Loshuertos and Enriquez, 2016). Recent work has also demonstrated that supercomplex assembly status also affects mitochondrial ROS production. For instance, O2˙/H2O2 production by complex I is higher when dissociated from complex III (Maranzana et al., 2013). Although little information exists on how supercomplex assemblies physically influence ROS production, it is possible that (1) dissociation of complex I from III makes FMN more accessible for electron transfer to O2 forming O2˙/H2O2 (Maranzana et al., 2013) and (2) FMN in complex I is more amenable to over reduction due to diminished electron transfer to complex III increasing electron availability for the monovalent reduction of O2. Thus, although speculative, it is likely that increased O2˙/H2O2 production by destabilized respirasomes is associated with increased electron availability and potentially increased access of O2 to electron donating sites.

Controlling mitochondrial ROS production by redox signals

Protein cysteine thiols can be subjected to a range of different redox modifications that depend on the concentration of H2O2, glutathione (GSH) and glutathione disulfide (GSSG), H2S, availability of nitric oxide (NO) and its adducts (like S-nitroso-glutathione or S-nitroso-coenzyme A), proximity of neighboring thiols allowing formation of intra- or inter-molecular disulfide bonds, or amides which can react with sulfenic acid to form sulfonamide. Most work in terms of understanding the impact of redox reactions on protein function has focused on the direct oxidation of cysteine thiols by H2O2 or NO. The direct oxidation of protein cysteine thiols by H2O2 is referred to as sulfenylation where a protein cysteine thiol (SH) is oxidized to a sulfenic acid (SOH). Notably, this reaction depends on the ionization of the sulfur group on cysteine forming a thiolate anion, a strong nucleophile that can attack H2O2. It has been postulated that H2O2 can negatively feedback on sites of production or on uncoupling proteins to control mitochondrial ROS release through the formation of SOH groups (Chouchani et al., 2016). However, very few proteins in mitochondria have been found to undergo sulfenylation. In addition, it was recently suggested that sulfenylation does not fit the criteria for regulation of proteins by covalent modification (Mailloux et al., 2016a,b). Indeed, to serve as an efficient regulatory system, the covalent modification needs to be specific, reversible, rapid, and fulfill some physiological role (Shelton et al., 2005). Hydrogen peroxide reacts slowly with cysteine residues and, to date, no known enzymes have been found to mediate these reactions nor do proteins display “sulfenylation motifs”. Moreover, SOH groups are highly unstable, reacting with various electrophiles forming different adducts (Nagy et al., 2007). A recent study showed that proteins are more likely to undergo S-glutathionylation and if a protein is sulfenylated it is rapidly resolved by neighboring amides to form sulfonamides (Forman et al., 2017). Sulfenylation can be reversed by the S-glutathionylation of SOH groups (Applegate et al., 2008). Sulfenic acid can also be oxidized by H2O2 forming sulfinic acid (SO2H), which can be reversed by sulfiredoxin (Rhee and Kil, 2016). However, these are protective mechanisms rather than regulatory modifications because both prevent the irreversible oxidation of thiols to sulfonic acid (SO3H). Collectively, it is unlikely that H2O2 directly regulates its own production in mitochondria as it does not fit the criteria of a signaling molecule. It is worthwhile noting that some proteins in the cell do exhibit sensitivity towards sulfenylation such as PRX proteins, which plays a role in either redox relay or floodgate signaling (discussed in detail in the section ‘Steroidogenesis in the adrenal cortex and circadian rhythms’). Modification of cysteine thiols by nitric oxide (S-nitrosylation) also suffers from the same issues, except a number of mitochondrial proteins have been found to undergo modification by NO. In addition, NO is well documented to serve as an inhibitor for complex IV of the respiratory chain. Thus, NO can still exert regulatory effects on the respiratory chain even though it still remains uncertain as to whether or not S-nitrosylation is a regulatory mechanism or a nonselective protein modification associated with cellular stress.

An emerging candidate for the post-translational control of proteins by redox signaling in the mitochondria is protein S-glutathionylation. This involves the conjugation and removal of a glutathione moiety from a protein. The first characteristic that sets S-glutathionylation apart from the other redox modifications is that it can proceed either spontaneously or enzymatically. The former is usually involved in protecting proteins during oxidative distress whereas the latter regulates protein function in response to normal physiological stimuli that alter local redox environments and buffering networks. The ratio of GSH/GSSG is usually kept quite high in the mitochondria, around 50–100, but this can be lowered in response to an increased H2O2 concentration (either higher than normal H2O2 formation by the respiratory chain or during the mitochondria-mediated quenching of cellular ROS). This increase in GSSG can induce the S-glutathionylation of proteins. Non-enzymatic S-glutathionylation has been documented to only occur when GSH/GSSG approaches 1 since some proteins have been found to have a KOX of ~1 (Gallogly and Mieyal, 2007). This occurs under oxidative distress when H2O2 is high and glutathione is actively being used for its degradation. Therefore, during oxidative distress proteins are S-glutathionylated to protect from irreversible damage (Dalle-Donne et al., 2009). Importantly, this modification is reversed by glutaredoxins (Grx; discussed in more detail below) which remove GSH and reactivate the protein. However, some proteins have a KOX with GSSG of ~50, meaning that under physiological conditions certain proteins can be S-glutathionylated spontaneously (Gallogly and Mieyal, 2007). In fact, some cellular proteins display basal S-glutathionylation under physiological conditions. For example, complex II (succinate dehydrogenase; SDH) has been found to be basally S-glutathionylated in heart tissue, which maintains its activity and diminishes the production of O2˙/H2O2 (Chen et al., 2007). 2-oxoglutarate dehydrogenase (OGDH) has also been found to undergo S-glutathionylation when GSH is abundant, a reaction thought to occur following the formation of a glutathionyl radical and its reaction with a protein cysteine thiol (Mailloux et al., 2016a,b). It has been suggested that, due to its highly folded nature, the mitochondrial inner membrane can generate highly variable redox micro-environments where the levels of GSH and GSSG can differ substantially (Drose et al., 2014). It is entirely plausible that this can occur when mitochondria transition between different states of respiration. For example, mitochondria display periodic changes in H2O2 and GSH levels when transitioning between state 3 and state 4 respiration (Cortassa et al., 2014). As mitochondria transition towards state 4 respiration, GSSG levels rise due to increased H2O2 production whereas induction of state 3 respiration has the opposite effect (Cortassa et al., 2014). This could have a strong influence on whether or not a mitochondrial protein can under spontaneous protein S-glutathionylation. Overall, nonspontaneous S-glutathionylation reactions are required to protect mitochondrial proteins, such as OGDH or complex I, from irreversible deactivation during oxidative distress. However, it may also be possible to spontaneously modify proteins with GSH if (1) the KOX is high enough and (2) the micro-environment GSH/GSSG gradient favors a spontaneous reaction.

Protein S-glutathionylation is also enzymatically mediated. These reactions are catalyzed by glutaredoxin (Grx), a small heat stable thiol oxidoreductase with a thioredoxin fold that mediates the conjugation and removal of glutathione from a target protein. GRX1 was the first thiol oxidoreductase found to catalyze the deglutathionylation of protein targets (Mannervik and Axelsson, 1980). GRX1 deglutathionylates protein cysteine thiols through a simple nucleophilic displacement mechanism involving thiol disulfide exchange between the active site cysteine in GRX1 and the target protein (Gallogly et al., 2009). It was later found to also catalyze the forward reaction, conjugating glutathione to a protein cysteine thiol. Unlike the deglutathionylation of target proteins, less is known about the mechanism for protein S-glutathionylation by GRX1. Although still speculative, it is possible that GRX1 mediates protein S-glutathionylation by first being S-glutathionylated itself allowing for transfer of glutathione to a protein target. Recent work has shown that GRX1 can be S-glutathionylated, indicating that this may account for how GRX1 can conjugate glutathione to cysteine (Dong et al., 2016). This can also be demonstrated with the cellular glutathione probe, GRX1-roGFP2, which changes fluorescence following its GSSG-mediated S-glutathionylation. GRX1 is also found in the intermembrane space of mitochondria where it has been shown to play a part in mitochondrial protein import (Pai et al., 2007). By contrast, glutaredoxin-2 (GRX2) is found in the matrix of mitochondria. Although GRX1 and GRX2 show very low homology with one another, both enzymes utilize a common mechanism for the reversible S-glutathionylation of proteins (Gallogly et al., 2009). It has been found that GRX2-mediated S-glutathionylation or deglutathionylation reactions are highly sensitive to changes in the surrounding redox environment. A high GSH/GSSG ratio activates the deglutathionylase activity of GRX2 whereas high H2O2 and a low GSH/GSSG redox pair have the opposite effect (Mailloux et al., 2014). GRX2 also forms an inactive dimer by coordinating an iron sulfur center with two GSH that can be disassembled by O2˙ yielding two active GRX2 monomers and two GSH (which are required to form the 2Fe-2S cluster) (Lillig et al., 2005). The importance of GRX2 in mediating the reversible S-glutathionylation of proteins is underscored by the consequences of its deletion (reviewed in detail in Mailloux, 2016). Briefly, loss of GRX2 can compromise mitochondrial respiration and ATP production and alter O2˙/H2O2 production, which is associated with the development of heart disease, hypertension, cataracts, and perturbations in embryonic development. In addition to the glutaredoxin enzymes, it has also been found that glutathione S-transferase can catalyze the reversible S-glutathionylation of cellular proteins (Klaus et al., 2013). Pi and Mu-members of the GST family have been found to catalyze the S-glutathionylation of AMP kinase, the master energy gauge for the cell (Klaus et al., 2013). It remains to be determined if GST isoforms are also required to catalyze these reactions in mitochondria. The deglutathionylation of target proteins has also been found to be quite rapid ranging from ~102 to 105m−1 s−1, which depends on the pKa of the target protein cysteine thiol and target proteins also have S-glutathionylation motifs, making this modification highly specific (Jensen et al., 2014). Thus, in contrast to sulfenylation or S-nitrosylation, protein S-glutathionylation reactions are rapid, reversible, highly specific, and sensitive to changes in H2O2 and the overall redox environment making this redox signal a strong candidate for regulating mitochondrial ROS production.

Protein S-glutathionylation regulates mitochondrial O2˙/H2O2 production

Regulation of complex I

One of the most extensively studied enzymes modified by S-glutathionylation is complex I of the electron transport chain. Complex I is composed of 45 subunits and catalyzes the two electron oxidation of NADH and reduction of ubiquinone to ubiquinol which drives the transfer of four protons from the matrix to the intermembrane space (Verkhovskaya and Bloch, 2013). This respiratory complex is also historically viewed as one of the most important sites for ROS production in mammalian mitochondria. Interest in whether or not complex I can be modulated by redox signals first arose when it was revealed that diethyl maleate and iodoacetamide, which specifically bind to protein cysteine thiols, alters its activity (Balijepalli et al., 1999). Intriguingly, these alterations correlated with a depletion of the mitochondrial glutathione pool (Balijepalli et al., 1999). Afterwards, Taylor and coworkers confirmed the presence of redox active thiols on complex I using membrane permeable iodobutyl triphenylphosphonium (IBTP) and various thiol modifying chemicals (diamide and GSSG) and reducing agents that can reverse protein S-glutathionylation (Taylor et al., 2003). It was also found that complex I underwent S-glutathionylation which increased O2˙ release from the respiratory enzyme (Taylor et al., 2003). Around the same period, GRX2 was characterized in mitochondria (Gladyshev et al., 2001) and it was thus hypothesized by the Murphy group that GRX2 was responsible for driving the S-glutathionylation of complex I. In the presence of oxidized glutathione and GRX2, complex I was found to be one of the most persistently S-glutathionylated proteins in mitochondrial membrane preparations (Beer et al., 2004) In addition, it was found by Beer and colleagues that S-glutathionylation of complex I by GRX2 was sensitive to changes in redox buffering capacity, and specifically by alterations in the GSH/GSSG ratio (Beer et al., 2004). Moreover, it has been found that the decrease in complex I activity following S-glutathionylation can be reversed by reducing agents like DTT, GSH, or purified GRX1 or GRX2 (Beer et al., 2004; Mailloux et al., 2014). The impact of S-glutathionylation on ROS formation by complex I has shown mixed results. An increase in ROS production has been shown by some groups (Taylor et al., 2003; Mailloux et al., 2014), while another study reported that S-glutathionylation decreases ROS production (Hurd et al., 2008). It is worth noting that the chief sites for S-glutathionylation are NDUSF1 and NDUFV1, which form part of the N-module of complex I (Mailloux et al., 2014). The N-module makes contact with the matrix and is thus the site for NADH oxidation. Modification of these two proteins has been proposed to limit access of electrons to the FMN site of complex I thus diminishing ROS production (Hurd et al., 2008). Modification of either protein, which lowers complex I activity, would also limit electron transfer to the ubiquinone-binding site, another putative ROS forming site. As for increasing O2˙/H2O2 production, another putative S-glutathionylation in complex I happens to be the ND3 subunit, which forms part of the ubiquinone binding pocket (Drose et al., 2016). Its S-glutathionylation could drive up O2˙/H2O2 formation by allowing the over-reduction of the FMN prosthetic group. Another possibility is that during long-term complex I S-glutathionylation, ROS production may be increased from other sites of formation due to NADH accumulation (e.g. via reverse electron flow through OGDH and PDH, discussed further below) (Mailloux et al., 2014). It had been proposed previously that complex I can adopt different S-glutathionylation states, which could influence O2˙/H2O2 formation under physiological and pathological conditions (Mailloux et al., 2014). This hypothesis was referred to as the ‘complex I cysteine code’ where, under physiological conditions, adoption of different S-glutathionylation states would ultimately influence how much ROS is emitted from mitochondria, which could affect redox signaling in mitochondria and throughout the cell. Conversely disruption of this so-called ‘cysteine code’ by disabling reversible S-glutathionylation reactions could have disastrous consequences due to long-term inhibition of complex I activity and uncontrolled O2˙/H2O2 production (either by complex I or other sites) (Mailloux et al., 2014).

2-Oxoglutarate dehydrogenase and pyruvate dehydrogenase

Complex I and III are historically viewed as the only sources of mitochondrial ROS. However, evidence collected over the past decade has shown that several other enzymes in mitochondria are also high capacity sites for O2˙/H2O2 formation. Indeed, it has been documented that skeletal muscle mitochondria can contain up to 12 ROS emitting sites. Of these sites it has been shown that 2-oxoglutarate dehydrogenase (OGDH) and pyruvate dehydrogenase (PDH) produce ~8× and ~4× more O2˙/H2O2 than complex I under experimental conditions where carbon is being oxidized by Krebs cycle enzymes (Quinlan et al., 2014). Similar observations have been made in liver and cardiac mitochondria where (1) OGDH and PDH are higher capacity sites for ROS formation than complex I and (2) OGDH is a more effective O2˙/H2O2 generator than PDH (Slade et al., 2017). It has also been found that purified OGDH and PDH both produce O2˙/H2O2 during reverse electron transfer (RET) from NADH (Ambrus et al., 2015; Mailloux et al., 2016a,b). Moreover, O2˙/H2O2 production during RET can occur at physiological NADH concentrations [0.1–1 μm, NADH/NAD+ ratio is 1/8 and concentration of NAD(H) is ~800 μm indicating physiological levels of NADH can stimulate ROS production by OGDH and PDH during reverse electron flow] and the rate of generation increases rapidly at higher NADH levels (Tretter and Adam-Vizi, 2004; Mailloux et al., 2016a,b). Overall, this would indicate that OGDH and PDH can form O2˙/H2O2 following reverse electron transfer from NADH, even during normal mitochondrial function. This would also indicate that both enzymes might be important sites of production under pathological conditions when NADH levels are increased due to a deficiency in complex I activity (Tretter and Adam-Vizi, 2005).

It has been known for some years that OGDH serves as a putative redox sensor (Gibson et al., 2005; McLain et al., 2011). Indeed, several studies over a decade ago had found that H2O2 can deactivate OGDH through oxidation of its lipoic residues (Tretter and Adam-Vizi, 1999). Afterwards it was uncovered that OGDH is also a source of O2˙/H2O2 leading to the postulate that H2O2 can feedback and limit its own production through the deactivation of the enzyme complex (McLain et al., 2011). Two studies showed that the reducing agent DTT or GRX1 could reactivate OGDH (Nulton-Persson et al., 2003; Applegate et al., 2008). Further inquiries into the modification of OGDH during oxidative stress showed that the lipoic acid residues on the E2 subunit were reversibly S-glutathionylated (Applegate et al., 2008; McLain et al., 2013). Considering that the vicinal lipoic acid thiols in OGDH are amenable to irreversible oxidation by H2O2, it was postulated that S-glutathionylation and its subsequent reversal was required to protect the enzyme complex from over-oxidation and reactivate it after the oxidative distress had subsided (McLain et al., 2011). It was also put forth that the S-glutathionylation of OGDH may also limit mitochondrial ROS production by diminishing NADH production and its subsequent oxidation by complex I (McLain et al., 2011). However, OGDH is also a major O2˙/H2O2 generator and thus its S-glutathionylation could also serve as a means of regulating H2O2 production. Our group recently showed that purified OGDH from porcine heart and OGDH in liver mitochondria can undergo S-glutathionylation, which is required to control O2˙/H2O2 release from the enzyme complex (Mailloux et al., 2016a,b; O’Brien et al., 2017). In two separate studies, it was found that the S-glutathionylation of the E2 subunit by either GSSG or diamide induces a significant decrease in O2˙/H2O2 release during 2-oxoglutarate oxidation (Mailloux et al., 2016a,b; O’Brien et al., 2017). Using purified OGDH, it was found that this decrease in O2˙/H2O2 production correlated strongly with a decrease in NADH production (Mailloux et al., 2016a,b). This would also diminish O2˙/H2O2 formation by the electron transport chain given that NADH production would be limited. Research into whether or not a high GSH/GSSG ratio or increased availability of GSH alone can affect ROS production led to the unique observation that reduced glutathione can actually augment O2˙/H2O2 release from OGDH through its S-glutathonylation. It was also found that GSH, at a higher concentration (>1 mm), could S-glutathionylate OGDH (Mailloux et al., 2016a,b). However, this event actually occurs on the E1 subunit, which augmented O2˙/H2O2 production by OGDH (Mailloux et al., 2016a,b). It should be noted here that OGDH produces O2˙/H2O2 from the E1 and E3 subunits; the former via thiamine radical formation and latter through its FAD prosthetic group (Nemeria et al., 2014). Thus, the GSH-mediated S-glutathionylation of the E1 subunit likely results in thiamine radical accumulation augmenting ROS production (Figure 3). The calculated EC50 for the GSH-mediated increase in ROS production by OGDH was found to be ~2.125 mm for ROS genesis and the IC50 for NADH production was ~2.819 mm (Mailloux et al., 2016a,b). In addition, purified GRX2 reversed these effects for both the purified form of OGDH and OGDH in permeabilized liver mitochondria. Therefore, when GSSG is high, it S-glutathionylates OGDH to limit O2˙/H2O2 production, which would be required to diminish production when ROS levels are high (Figure 3). Conversely, high GSH has the opposite effect, augmenting O2˙/H2O2 production when mitochondrial redox buffering capacity is reduced enough to control higher than normal ROS formation (Figure 4). Collectively, OGDH can serve as an overall redox buffering sensor, increasing and decreasing its ROS production rate in response to fluctuations in GSH and GSSG availability, which could influence the capacity of mitochondria to use H2O2 as a secondary messenger.

Figure 3: GSH and GSSG have different effects on ROS release from 2-oxoglutarate dehydrogenase.(A) At low ROS level, 2-oxoglutarate dehydrogenase (OGDH) electrons yielded from the combustion of 2-oxoglutarate are used to drive NADH and ROS formation. When H2O2 levels are higher than normal or a mitochondrion is experiencing mild oxidative distress, GSSG levels increase, resulting in the S-glutathionylation of the vicinal thiols on the E2 subunit of OGDH blocking electron transfer to the E3 subunit. This decreases ROS production by OGDH whilst simultaneously protecting the enzyme from irreversible oxidation. S-glutathionylation also decreases NADH production, which would lower ROS formation by the respiratory complexes. (B) At high enough levels, GSH can S-glutathionylate the E1 subunit of OGDH amplifying ROS production. The higher GSH levels can allow for easy buffering of any excess ROS formed during substrate oxidation. In addition, high GSH can maintain GRX2 in its inactive dimeric form. Activation of GRX2 by high ROS results in the deglutathionylation of OGDH and the restoration of NADH production.
Figure 3:

GSH and GSSG have different effects on ROS release from 2-oxoglutarate dehydrogenase.

(A) At low ROS level, 2-oxoglutarate dehydrogenase (OGDH) electrons yielded from the combustion of 2-oxoglutarate are used to drive NADH and ROS formation. When H2O2 levels are higher than normal or a mitochondrion is experiencing mild oxidative distress, GSSG levels increase, resulting in the S-glutathionylation of the vicinal thiols on the E2 subunit of OGDH blocking electron transfer to the E3 subunit. This decreases ROS production by OGDH whilst simultaneously protecting the enzyme from irreversible oxidation. S-glutathionylation also decreases NADH production, which would lower ROS formation by the respiratory complexes. (B) At high enough levels, GSH can S-glutathionylate the E1 subunit of OGDH amplifying ROS production. The higher GSH levels can allow for easy buffering of any excess ROS formed during substrate oxidation. In addition, high GSH can maintain GRX2 in its inactive dimeric form. Activation of GRX2 by high ROS results in the deglutathionylation of OGDH and the restoration of NADH production.

Figure 4: S-glutathionylation is required to control ROS production by pyruvate dehydrogenase.(A) During forward electron flow from pyruvate to NAD+, pyruvate dehydrogenase (PDH) forms high amounts of ROS, which can be inhibited by the S-glutathionylation of the vicinal thiols on the lipoic acid in the E2 subunit. ROS emitted from PDH may be required for cell signaling and thus S-glutathionylation is required to desensitize the H2O2 signal. (B) PDH can form ROS following reverse electron transfer from NADH. During oxidative distress GSSG accumulates resulting in the spontaneous S-glutathionylation of PDH, which augments ROS production during reverse electron transfer. Oxidative distress is also associated with the deactivation of complex I, resulting in NADH accumulation which is then oxidized by PDH. High ROS production due to reverse electron flow when NADH is high and PDH is S-glutathionylated, results in GRX2 activation and the deglutathionylation of PDH. This restores PDH activity and lowers ROS production by reverse electron flow from NADH.
Figure 4:

S-glutathionylation is required to control ROS production by pyruvate dehydrogenase.

(A) During forward electron flow from pyruvate to NAD+, pyruvate dehydrogenase (PDH) forms high amounts of ROS, which can be inhibited by the S-glutathionylation of the vicinal thiols on the lipoic acid in the E2 subunit. ROS emitted from PDH may be required for cell signaling and thus S-glutathionylation is required to desensitize the H2O2 signal. (B) PDH can form ROS following reverse electron transfer from NADH. During oxidative distress GSSG accumulates resulting in the spontaneous S-glutathionylation of PDH, which augments ROS production during reverse electron transfer. Oxidative distress is also associated with the deactivation of complex I, resulting in NADH accumulation which is then oxidized by PDH. High ROS production due to reverse electron flow when NADH is high and PDH is S-glutathionylated, results in GRX2 activation and the deglutathionylation of PDH. This restores PDH activity and lowers ROS production by reverse electron flow from NADH.

In regard to redox sensing less is known about PDH. However, recent work by our group and others found that it serves as a GSH and GSSG sensor much like OGDH (O’Brien et al., 2017). Using purified PDH and permeabilized muscle fibers, Fisher-Wellman and coworkers found that depletion of glutathione pools or inhibition of NADPH production augmented pyruvate-driven O2˙/H2O2 production (Fisher-Wellman et al., 2013, 2015). It was then postulated by the same group that PDH was able to sense changes in mitochondrial redox buffering capacity, potentially through S-glutathionylation, which alters ROS production (Fisher-Wellman et al., 2013). However, based on the experimental design it was unclear whether or not PDH was actually sensing any changes in mitochondrial redox buffering capacity or if the increase was simply due to the CDNB-mediated depletion of glutathione pools. Our group recently found that PDH is S-glutathionylated on its E2 subunit and to a lesser extent its E1 and E3 subunits (O’Brien et al., 2017). Gergondey and colleagues also recently showed that H2O2 and diamide can drive the S-glutathionylation of PDH (Gergondey et al., 2017). During pyruvate oxidation, it was found that the diamide or disulfiram mediated S-glutathionylation of purified PDH or PDH in liver mitochondria decreased O2˙/H2O2 production (O’Brien et al., 2017). This was also associated with a decrease in NADH formation which correlated with the S-glutathionylation of the E2 subunit of PDH. It was also found that the GSSG-mediated S-glutathionylation of purified PDH augmented ROS production during RET which could be reversed by GRX2 (O’Brien et al., 2017). Therefore, the S-glutathionylation of PDH serves as a mechanism for sensing changes in mitochondrial redox buffering capacity, which can sharply alter how much O2˙/H2O2 is formed by the enzyme complex. During forward electron flow from pyruvate to NAD+, S-glutathionylation of the E2 subunit is required to lower ROS production (Figure 4). Moreover, S-glutathionylation of PDH almost completely abolished NADH production, which would limit O2˙/H2O2 production by the electron transport chain. Overall, this could play an integral role in influencing mitochondrial H2O2 release and signaling to the rest of the cell given that PDH is a vital source of O2˙/H2O2 and an important entry point into the Krebs cycle for carbon yielded from glucose metabolism (Figure 4). In regard to production during RET, we postulate this may reflect pathological conditions as S-glutathionylation of PDH augmented NADH driven ROS production by eight times (O’Brien et al., 2017). This could contribute to oxidative distress when GSSG levels are high enough and NADH turnover by complex I is limiting, for example in dysfunctional mitochondria (Figure 4). Since GRX2 can reverse the GSSG-mediated S-glutathionylation of PDH and the sharp increase in O2˙/H2O2 production, it is likely that it is required to protect mitochondria from any potential oxidative distress associated with the over production of ROS (Figure 4).

Succinate dehydrogenase

Succinate dehydrogenase (SDH), also known as complex II of the electron transport chain, has also been identified as a significant source of O2˙/H2O2 in the mitochondria. SDH can generate O2˙/H2O2 when electrons flow in both the forward and reverse directions, and production has been pinpointed to the flavin site of the enzyme (Quinlan et al., 2012; Drose, 2013). To function properly, it was found that complex II in cardiac mitochondria is basally S-glutathionylated on Cys90 of the SDHA subunit that houses its FAD prosthetic group (Chen et al., 2007). This was reported to maintain low ROS production by complex II. This also maintains optimal activity for complex II allowing for efficient oxidation of succinate and the subsequent transfer of electrons to the ubiquinone pool (Chen et al., 2007). However, during ischemia-reperfusion injury to the myocardium, the SDHA subunit is deglutathionylated by an unknown mechanism resulting in higher than normal O2˙/H2O2 production, diminished electron flow to the ubiquinone pool, and oxidative damage to cardiac tissue (Chen et al., 2007). It has been postulated that S-glutathionylation of Cys90 induces conformational change in the SDHA subunit leading to increased electron movement in complex II limiting O2˙/H2O2 formation by FAD whilst maintaining high enzyme activity (Chen et al., 2007). Therefore, S-glutathionylation also plays a vital role in regulating O2˙/H2O2 production from complex II, another putative high capacity site in mitochondria. However, under ischemic-reperfusion conditions other enzymes like complex I, PDH, and OGDH might also display different S-glutathionylation profiles which may favor higher than normal O2˙/H2O2 production. For example, during ischemia, complex I transitions from an activate A-state to a deactivate D-state where a cysteine residue on the ND3 subunit becomes amenable to oxidative modification (Drose et al., 2016). This allows NADH to accumulate in the matrix of mitochondria. Upon the reintroduction of O2 (reperfusion), complex I slowly transitions back to its A-form. However, oxidative modification of the ND3 subunit, for instance by S-glutathionylation, can prolong complex I deactivation following reperfusion. It is possible that complex I is maintained in the D-state through the S-glutathionylation of its ND3 subunit prompting prolonged NADH accumulation, which can drive O2˙/H2O2 production via RET through OGDH and PDH. This concept is currently being tested by our group with a complex I deficient mouse model.

Uncoupling proteins-2 and 3

As mentioned at the beginning of this section, UCP2 and UCP3 have been implicated in controlling ROS production in mitochondria by increasing proton return to the matrix. However, it was also iterated that this function is still controversial as other groups have found that these integral membrane proteins can translocate other solutes. The putative role of uncoupling proteins in controlling mitochondrial ROS production through the induction of leaks was first suggested for UCP2, where it was found to play a role in regulating H2O2 levels (Negre-Salvayre et al., 1997). Later, it was found UCP1-3 were also able to decrease mitochondrial O2˙/H2O2 release via a negative feedback loop involving the O2˙ or 4-hydroxy-2-nonenal mediated activation of proton return (Echtay et al., 2002, 2003). However, numerous studies failed to reproduce these effects. This is likely associated with several factors, including (1) 4-hydroxy-2-nonenal does not selectively bind uncoupling proteins and has been found to induce leaks through ANT and the mitochondrial permeability transition pore and (2) O2˙ is membrane impermeant and a weak oxidant and reductant and thus reacts with very few biological molecules including proteins. However, this does not necessarily mean other factors that alter redox buffering capacity do not influence uncoupling protein activity. For example, Chouchani and colleagues recently found that UCP1 can be activated via the H2O2 driven oxidation of cysteine residues that make contact with the matrix (Chouchani et al., 2016). Similar observations have been made with UCP3, which has been found to be S-glutathionylated on Cys25 and Cys259 (Mailloux et al., 2011). S-glutathionylation of UCP3 causes a decrease in state 4 respiration in primary cells and skeletal muscle mitochondria (Mailloux et al., 2011). S-glutathionylation of UCP3 was also discovered to be catalyzed by the enzyme GRX2 (Mailloux et al., 2013). In GRX2 knock out mice, UCP3 was less glutathionylated which resulted in increased state 4 respiration (Mailloux et al., 2013). UCP2 has also been found to be S-glutathionylated, which was shown to play a part in regulating glucose-stimulated insulin release (Mailloux et al., 2012ab). In regard to O2˙/H2O2 changes in the S-glutathionylation state of UCP2 and UCP3 induces small but significant changes in mitochondrial O2˙/H2O2 production (Mailloux et al., 2011). Whether or not this is associated with changes in proton return to the matrix still requires clarification.

Conclusion

The formation of H2O2 by mitochondria was identified almost 50 years ago (Loschen et al., 1971). Once considered a major source for a mammalian cell’s oxidative burden, it is now appreciated that O2˙/H2O2 release from mitochondria regulates cellular functions. The emergence of new technologies that selectively detect changes in H2O2 levels in live cells and tissues have allowed for the characterization of new signaling networks between mitochondria and the rest of the cell. Over the past decade, it has become apparent that mitochondrial H2O2 regulates cell proliferation, stress signaling, insulin release and sensitivity, steroidogenesis, T-cell activation, and cell differentiation. Since H2O2 can be damaging at high levels it is important to control its production and overall concentration. Antioxidant systems are integral for removing ROS but mitochondria also rely on regulatory mechanisms to control its formation. Proton leaks have often been viewed as one of the main mechanisms for controlling mitochondrial H2O2 release. However, it has become evident that the rate of electron leak and O2˙/H2O2 formation does not necessarily correlate with the rate of proton return and respiration. Thus, other mechanisms are utilized to modulate O2˙/H2O2 production from individual sites directly. This includes redox signals like S-glutathionylation which can feedback on sites of production to limit O2˙/H2O2 formation. This is achieved by blocking electron flow to different sites of production which lowers O2˙/H2O2 production or by augmenting electron transfer limiting the number of electrons available at sites production. The S-glutathionylation of proteins is also likely associated with H2O2 and NADPH mediated alterations in the redox state of glutathione pool; H2O2 promotes S-glutathionylation by oxidizing glutathione while provision of NADPH has the opposite effect. This could potentially serve as the interface required for adjusting cell functions and mitochondrial metabolism in response to changes in the intra- and extracellular environment. It is also likely that S-glutathionylation is required to safeguard cells from the potentially harmful effects of ROS. Indeed, deregulation of these pathways is associated with the development of different pathologies which is related to high mitochondrial ROS production. The role of mitochondrial redox buffering networks in regulating ROS release is still a novel concept and whether it plays a part in regulating the mitokine function of H2O2 still remains unexplored. However, based on the available evidence it would seem that S-glutathionylation reactions are a strong candidate for the regulation of H2O2 messaging from mitochondria to the rest of the cell.

Acknowledgements

This work is dedicated to the memory of Jo-Ann Mailloux. Funding was provided by the Natural Sciences and Engineering Research Council of Canada (NSERC # RGPIN-2016-04829). N.K. was funded by the Canadian Queen Elizabeth II Diamond Jubilee Scholarship Program.

References

Abidi, P., Zhang, H., Zaidi, S.M., Shen, W.J., Leers-Sucheta, S., Cortez, Y., Han, J., and Azhar, S. (2008). Oxidative stress-induced inhibition of adrenal steroidogenesis requires participation of p38 mitogen-activated protein kinase signaling pathway. J. Endocrinol. 198, 193–207.10.1677/JOE-07-0570Search in Google Scholar

Acin-Perez, R., Fernandez-Silva, P., Peleato, M.L., Perez-Martos, A., and Enriquez, J.A. (2008). Respiratory active mitochondrial supercomplexes. Mol. Cell 32, 529–539.10.1016/j.molcel.2008.10.021Search in Google Scholar

Affourtit, C., Jastroch, M., and Brand, M.D. (2011). Uncoupling protein-2 attenuates glucose-stimulated insulin secretion in INS-1E insulinoma cells by lowering mitochondrial reactive oxygen species. Free Radic. Biol. Med. 50, 609–616.10.1016/j.freeradbiomed.2010.12.020Search in Google Scholar

Allister, E.M., Robson-Doucette, C.A., Prentice, K.J., Hardy, A.B., Sultan, S., Gaisano, H.Y., Kong, D., Gilon, P., Herrera, P.L., Lowell, B.B., et al. (2013). UCP2 regulates the glucagon response to fasting and starvation. Diabetes 62, 1623–1633.10.2337/db12-0981Search in Google Scholar

Ambrus, A., Nemeria, N.S., Torocsik, B., Tretter, L., Nilsson, M., Jordan, F., and Adam-Vizi, V. (2015). Formation of reactive oxygen species by human and bacterial pyruvate and 2-oxoglutarate dehydrogenase multienzyme complexes reconstituted from recombinant components. Free Radic. Biol. Med. 89, 642–650.10.1016/j.freeradbiomed.2015.10.001Search in Google Scholar

Andrikopoulos, S. (2016). Chewing the fat for better insulin secretion. Mol. Metab. 5, 3–4.10.1016/j.molmet.2015.10.003Search in Google Scholar

Anedda, A., Rial, E., and Gonzalez-Barroso, M.M. (2008). Metformin induces oxidative stress in white adipocytes and raises uncoupling protein 2 levels. J. Endocrinol. 199, 33–40.10.1677/JOE-08-0278Search in Google Scholar

Applegate, M.A., Humphries, K.M., and Szweda, L.I. (2008). Reversible inhibition of α-ketoglutarate dehydrogenase by hydrogen peroxide: glutathionylation and protection of lipoic acid. Biochemistry 47, 473–478.10.1021/bi7017464Search in Google Scholar

Ardanaz, N., Yang, X.P., Cifuentes, M.E., Haurani, M.J., Jackson, K.W., Liao, T.D., Carretero, O.A., and Pagano, P.J. (2010). Lack of glutathione peroxidase 1 accelerates cardiac-specific hypertrophy and dysfunction in angiotensin II hypertension. Hypertension 55, 116–123.10.1161/HYPERTENSIONAHA.109.135715Search in Google Scholar

Balijepalli, S., Annepu, J., Boyd, M.R., and Ravindranath, V. (1999). Effect of thiol modification on brain mitochondrial complex I activity. Neurosci. Lett. 272, 203–206.10.1016/S0304-3940(99)00593-5Search in Google Scholar

Beer, S.M., Taylor, E.R., Brown, S.E., Dahm, C.C., Costa, N.J., Runswick, M.J., and Murphy, M.P. (2004). Glutaredoxin 2 catalyzes the reversible oxidation and glutathionylation of mitochondrial membrane thiol proteins: implications for mitochondrial redox regulation and antioxidant DEFENSE. J. Biol. Chem. 279, 47939–47951.10.1074/jbc.M408011200Search in Google Scholar PubMed

Bell, E.L., Klimova, T.A., Eisenbart, J., Moraes, C.T., Murphy, M.P., Budinger, G.R., and Chandel, N.S. (2007). The Qo site of the mitochondrial complex III is required for the transduction of hypoxic signaling via reactive oxygen species production. J. Cell. Biol. 177, 1029–1036.10.1083/jcb.200609074Search in Google Scholar PubMed PubMed Central

Bernard, J., Gee, J.B., and Khandwala, A.S. (1976). Oxygen metabolism in the alveolar macrophage: friend and foe? J. Reticuloendothel. Soc. 19, 229–236.Search in Google Scholar

Brand, M.D. (2016). Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radic. Biol. Med. 100, 14–31.10.1016/j.freeradbiomed.2016.04.001Search in Google Scholar PubMed

Brand, M.D. and Nicholls, D.G. (2011). Assessing mitochondrial dysfunction in cells. Biochem. J. 435, 297–312.10.1042/BJ20110162Search in Google Scholar PubMed PubMed Central

Chandel, N.S., Maltepe, E., Goldwasser, E., Mathieu, C.E., Simon, M.C., and Schumacker, P.T. (1998). Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc. Natl. Acad. Sci. USA 95, 11715–11720.10.1073/pnas.95.20.11715Search in Google Scholar PubMed PubMed Central

Chaudhuri, L., Srivastava, R.K., Kos, F., and Shrikant, P.A. (2016). Uncoupling protein 2 regulates metabolic reprogramming and fate of antigen-stimulated CD8+ T cells. Cancer Immunol. Immunother. 65, 869–874.10.1007/s00262-016-1851-4Search in Google Scholar PubMed PubMed Central

Chen, Y.R., Chen, C.L., Pfeiffer, D.R., and Zweier, J.L. (2007). Mitochondrial complex II in the post-ischemic heart: oxidative injury and the role of protein S-glutathionylation. J. Biol. Chem. 282, 32640–32654.10.1074/jbc.M702294200Search in Google Scholar PubMed

Chouchani, E.T., Kazak, L., Jedrychowski, M.P., Lu, G.Z., Erickson, B.K., Szpyt, J., Pierce, K.A., Laznik-Bogoslavski, D., Vetrivelan, R., Clish, C.B., et al. (2016). Mitochondrial ROS regulate thermogenic energy expenditure and sulfenylation of UCP1. Nature 532, 112–116.10.1038/nature17399Search in Google Scholar PubMed PubMed Central

Cortassa, S., O’Rourke, B., and Aon, M.A. (2014). Redox-optimized ROS balance and the relationship between mitochondrial respiration and ROS. Biochim. Biophys. Acta 1837, 287–295.10.1016/j.bbabio.2013.11.007Search in Google Scholar PubMed PubMed Central

Cox, A.G., Pearson, A.G., Pullar, J.M., Jonsson, T.J., Lowther, W.T., Winterbourn, C.C., and Hampton, M.B. (2009). Mitochondrial peroxiredoxin 3 is more resilient to hyperoxidation than cytoplasmic peroxiredoxins. Biochem. J. 421, 51–58.10.1042/BJ20090242Search in Google Scholar PubMed PubMed Central

Dalle-Donne, I., Rossi, R., Colombo, G., Giustarini, D., and Milzani, A. (2009). Protein S-glutathionylation: a regulatory device from bacteria to humans. Trends Biochem. Sci. 34, 85–96.10.1016/j.tibs.2008.11.002Search in Google Scholar PubMed

Davies, K.J. (2016). The oxygen paradox, oxidative stress, and ageing. Arch. Biochem. Biophys. 595, 28–32.10.1016/j.abb.2015.11.015Search in Google Scholar PubMed PubMed Central

Deponte, M. (2013). Glutathione catalysis and the reaction mechanisms of glutathione-dependent enzymes. Biochim. Biophys. Acta 1830, 3217–3266.10.1016/j.bbagen.2012.09.018Search in Google Scholar PubMed

Devadas, S., Zaritskaya, L., Rhee, S.G., Oberley, L., and Williams, M.S. (2002). Discrete generation of superoxide and hydrogen peroxide by T cell receptor stimulation: selective regulation of mitogen-activated protein kinase activation and fas ligand expression. J. Exp. Med. 195, 59–70.10.1084/jem.20010659Search in Google Scholar PubMed PubMed Central

Diemer, T., Allen, J.A., Hales, K.H., and Hales, D.B. (2003). Reactive oxygen disrupts mitochondria in MA-10 tumor Leydig cells and inhibits steroidogenic acute regulatory (StAR) protein and steroidogenesis. Endocrinology 144, 2882–2891.10.1210/en.2002-0090Search in Google Scholar PubMed

Dinkova-Kostova, A.T., and Abramov, A.Y. (2015). The emerging role of Nrf2 in mitochondrial function. Free Radic. Biol. Med. 88, 179–188.10.1016/j.freeradbiomed.2015.04.036Search in Google Scholar PubMed PubMed Central

Divakaruni, A.S. and Brand, M.D. (2011). The regulation and physiology of mitochondrial proton leak. Physiology (Bethesda) 26, 192–205.10.1152/physiol.00046.2010Search in Google Scholar PubMed

Dong, K., Wu, M., Liu, X., Huang, Y., Zhang, D., Wang, Y., Yan, L.J., and Shi, D. (2016). Glutaredoxins concomitant with optimal ROS activate AMPK through S-glutathionylation to improve glucose metabolism in type 2 diabetes. Free Radic. Biol. Med. 101, 334–347.10.1016/j.freeradbiomed.2016.10.007Search in Google Scholar PubMed

Drechsel, D.A. and Patel, M. (2010). Respiration-dependent H2O2 removal in brain mitochondria via the thioredoxin/peroxiredoxin system. J. Biol. Chem. 285, 27850–27858.10.1074/jbc.M110.101196Search in Google Scholar PubMed PubMed Central

Drose, S. (2013). Differential effects of complex II on mitochondrial ROS production and their relation to cardioprotective pre- and postconditioning. Biochim. Biophys. Acta 1827, 578–587.10.1016/j.bbabio.2013.01.004Search in Google Scholar PubMed

Drose, S., Brandt, U., and Wittig, I. (2014). Mitochondrial respiratory chain complexes as sources and targets of thiol-based redox-regulation. Biochim. Biophys. Acta 1844, 1344–1354.10.1016/j.bbapap.2014.02.006Search in Google Scholar PubMed

Drose, S., Stepanova, A., and Galkin, A. (2016). Ischemic A/D transition of mitochondrial complex I and its role in ROS generation. Biochim. Biophys. Acta 1857, 946–957.10.1016/j.bbabio.2015.12.013Search in Google Scholar PubMed PubMed Central

Duranteau, J., Chandel, N.S., Kulisz, A., Shao, Z., and Schumacker, P.T. (1998). Intracellular signaling by reactive oxygen species during hypoxia in cardiomyocytes. J. Biol. Chem. 273, 11619–11624.10.1074/jbc.273.19.11619Search in Google Scholar PubMed

Echtay, K.S., Roussel, D., St-Pierre, J., Jekabsons, M.B., Cadenas, S., Stuart, J.A., Harper, J.A., Roebuck, S.J., Morrison, A., Pickering, S., et al. (2002). Superoxide activates mitochondrial uncoupling proteins. Nature 415, 96–99.10.1038/415096aSearch in Google Scholar PubMed

Echtay, K.S., Esteves, T.C., Pakay, J.L., Jekabsons, M.B., Lambert, A.J., Portero-Otin, M., Pamplona, R., Vidal-Puig, A.J., Wang, S., Roebuck, S.J., et al. (2003). A signalling role for 4-hydroxy-2-nonenal in regulation of mitochondrial uncoupling. EMBO J. 22, 4103–4110.10.1093/emboj/cdg412Search in Google Scholar PubMed PubMed Central

Enriquez, J.A. (2016). Supramolecular organization of respiratory complexes. Annu. Rev. Physiol. 78, 533–561.10.1146/annurev-physiol-021115-105031Search in Google Scholar PubMed

Fisher-Wellman, K.H., Gilliam, L.A., Lin, C.T., Cathey, B.L., Lark, D.S., and Neufer, P.D. (2013). Mitochondrial glutathione depletion reveals a novel role for the pyruvate dehydrogenase complex as a key H2O2-emitting source under conditions of nutrient overload. Free Radic. Biol. Med. 65, 1201–1208.10.1016/j.freeradbiomed.2013.09.008Search in Google Scholar PubMed PubMed Central

Fisher-Wellman, K.H., Lin, C.T., Ryan, T.E., Reese, L.R., Gilliam, L.A., Cathey, B.L., Lark, D.S., Smith, C.D., Muoio, D.M., and Neufer, P.D. (2015). Pyruvate dehydrogenase complex and nicotinamide nucleotide transhydrogenase constitute an energy-consuming redox circuit. Biochem. J. 467, 271–280.10.1042/BJ20141447Search in Google Scholar PubMed PubMed Central

Forman, H.J., Davies, M.J., Kramer, A.C., Miotto, G., Zaccarin, M., Zhang, H., and Ursini, F. (2017). Protein cysteine oxidation in redox signaling: caveats on sulfenic acid detection and quantification. Arch. Biochem. Biophys. 617, 26–37.10.1016/j.abb.2016.09.013Search in Google Scholar PubMed PubMed Central

Gallogly, M.M. and Mieyal, J.J. (2007). Mechanisms of reversible protein glutathionylation in redox signaling and oxidative stress. Curr. Opin. Pharmacol. 7, 381–391.10.1016/j.coph.2007.06.003Search in Google Scholar PubMed

Gallogly, M.M., Starke, D.W., and Mieyal, J.J. (2009). Mechanistic and kinetic details of catalysis of thiol-disulfide exchange by glutaredoxins and potential mechanisms of regulation. Antioxid. Redox Signal. 11, 1059–1081.10.1089/ars.2008.2291Search in Google Scholar

Gergondey, R., Garcia, C., Marchand, C.H., Lemaire, S.D., Camadro, J.M., and Auchere, F. (2017). Modulation of the specific glutathionylation of mitochondrial proteins in the yeast Saccharomyces cerevisiae under basal and stress conditions. Biochem. J. 474, 1175–1193.10.1042/BCJ20160927Search in Google Scholar

Gibson, G.E., Blass, J.P., Beal, M.F., and Bunik, V. (2005). The α-ketoglutarate-dehydrogenase complex: a mediator between mitochondria and oxidative stress in neurodegeneration. Mol. Neurobiol. 31, 43–63.10.1385/MN:31:1-3:043Search in Google Scholar

Gladyshev, V.N., Liu, A., Novoselov, S.V., Krysan, K., Sun, Q.A., Kryukov, V.M., Kryukov, G.V., and Lou, M.F. (2001). Identification and characterization of a new mammalian glutaredoxin (thioltransferase), Grx2. J. Biol. Chem. 276, 30374–30380.10.1074/jbc.M100020200Search in Google Scholar PubMed

Go, Y.M., Chandler, J.D., and Jones, D.P. (2015). The cysteine proteome. Free Radic. Biol. Med. 84, 227–245.10.1016/j.freeradbiomed.2015.03.022Search in Google Scholar PubMed PubMed Central

Graham, K.A., Kulawiec, M., Owens, K.M., Li, X., Desouki, M.M., Chandra, D., and Singh, K.K. (2010). NADPH oxidase 4 is an oncoprotein localized to mitochondria. Cancer Biol. Ther. 10, 223–231.10.4161/cbt.10.3.12207Search in Google Scholar PubMed PubMed Central

Halliwell, B. (1996). Antioxidants in human health and disease. Annu. Rev. Nutr. 16, 33–50.10.1146/annurev.nu.16.070196.000341Search in Google Scholar PubMed

Hamilton, R.T., Walsh, M.E., and Van Remmen, H. (2012). Mouse models of oxidative stress indicate a role for modulating healthy aging. J. Clin. Exp. Pathol. (Suppl 4, 005).10.4172/2161-0681.S4-005Search in Google Scholar PubMed PubMed Central

Harper, M.E., Green, K., and Brand, M.D. (2008). The efficiency of cellular energy transduction and its implications for obesity. Annu. Rev. Nutr. 28, 13–33.10.1146/annurev.nutr.28.061807.155357Search in Google Scholar PubMed

Hurd, T.R., Requejo, R., Filipovska, A., Brown, S., Prime, T.A., Robinson, A.J., Fearnley, I.M., and Murphy, M.P. (2008). Complex I within oxidatively stressed bovine heart mitochondria is glutathionylated on Cys-531 and Cys-704 of the 75-kDa subunit: potential role of CYS residues in decreasing oxidative damage. J. Biol. Chem. 283, 24801–24815.10.1074/jbc.M803432200Search in Google Scholar PubMed PubMed Central

Imhoff, B.R. and Hansen, J.M. (2010). Extracellular redox environments regulate adipocyte differentiation. Differentiation 80, 31–39.10.1016/j.diff.2010.04.005Search in Google Scholar PubMed

James, A.M., Collins, Y., Logan, A., and Murphy, M.P. (2012). Mitochondrial oxidative stress and the metabolic syndrome. Trends Endocrinol. Metab. 23, 429–434.10.1016/j.tem.2012.06.008Search in Google Scholar PubMed

Jarvis, R.M., Hughes, S.M., and Ledgerwood, E.C. (2012). Peroxiredoxin 1 functions as a signal peroxidase to receive, transduce, and transmit peroxide signals in mammalian cells. Free Radic. Biol. Med. 53, 1522–1530.10.1016/j.freeradbiomed.2012.08.001Search in Google Scholar PubMed

Jastroch, M., Divakaruni, A.S., Mookerjee, S., Treberg, J.R., and Brand, M.D. (2010). Mitochondrial proton and electron leaks. Essays Biochem. 47, 53–67.10.1042/bse0470053Search in Google Scholar PubMed PubMed Central

Jensen, M.V., Joseph, J.W., Ronnebaum, S.M., Burgess, S.C., Sherry, A.D., and Newgard, C.B. (2008). Metabolic cycling in control of glucose-stimulated insulin secretion. Am. J. Physiol. Endocrinol. Metab. 295, E1287–E1297.10.1152/ajpendo.90604.2008Search in Google Scholar PubMed PubMed Central

Jensen, K.S., Pedersen, J.T., Winther, J.R., and Teilum, K. (2014). The pKa value and accessibility of cysteine residues are key determinants for protein substrate discrimination by glutaredoxin. Biochemistry 53, 2533–2540.10.1021/bi4016633Search in Google Scholar PubMed

Jones, D.P. (2015). Redox theory of aging. Redox Biol. 5, 71–79.10.1016/j.redox.2015.03.004Search in Google Scholar PubMed PubMed Central

Jones, D.P. and Sies, H. (2015). The redox code. Antioxid. Redox Signal. 23, 734–746.10.1089/ars.2015.6247Search in Google Scholar PubMed PubMed Central

Kaminski, M.M., Sauer, S.W., Klemke, C.D., Suss, D., Okun, J.G., Krammer, P.H., and Gulow, K. (2010). Mitochondrial reactive oxygen species control T cell activation by regulating IL-2 and IL-4 expression: mechanism of ciprofloxacin-mediated immunosuppression. J. Immunol. 184, 4827–4841.10.4049/jimmunol.0901662Search in Google Scholar PubMed

Kil, I.S., Lee, S.K., Ryu, K.W., Woo, H.A., Hu, M.C., Bae, S.H., and Rhee, S.G. (2012). Feedback control of adrenal steroidogenesis via H2O2-dependent, reversible inactivation of peroxiredoxin III in mitochondria. Mol. Cell 46, 584–594.10.1016/j.molcel.2012.05.030Search in Google Scholar PubMed

Klaus, A., Zorman, S., Berthier, A., Polge, C., Ramirez, S., Michelland, S., Seve, M., Vertommen, D., Rider, M., Lentze, N., et al. (2013). Glutathione S-transferases interact with AMP-activated protein kinase: evidence for S-glutathionylation and activation in vitro. PLoS One 8, e62497.10.1371/journal.pone.0062497Search in Google Scholar

Korshunov, S.S., Skulachev, V.P., and Starkov, A.A. (1997). High protonic potential actuates a mechanism of production of reactive oxygen species in mitochondria. FEBS Lett. 416, 15–18.10.1016/S0014-5793(97)01159-9Search in Google Scholar

Leloup, C., Tourrel-Cuzin, C., Magnan, C., Karaca, M., Castel, J., Carneiro, L., Colombani, A.L., Ktorza, A., Casteilla, L., and Penicaud, L. (2009). Mitochondrial reactive oxygen species are obligatory signals for glucose-induced insulin secretion. Diabetes 58, 673–681.10.2337/db07-1056Search in Google Scholar

Lillig, C.H., Berndt, C., Vergnolle, O., Lonn, M.E., Hudemann, C., Bill, E., and Holmgren, A. (2005). Characterization of human glutaredoxin 2 as iron-sulfur protein: a possible role as redox sensor. Proc. Natl. Acad. Sci. USA 102, 8168–8173.10.1073/pnas.0500735102Search in Google Scholar

Lombardi, A., Busiello, R.A., Napolitano, L., Cioffi, F., Moreno, M., de Lange, P., Silvestri, E., Lanni, A., and Goglia, F. (2010). UCP3 translocates lipid hydroperoxide and mediates lipid hydroperoxide-dependent mitochondrial uncoupling. J. Biol. Chem. 285, 16599–16605.10.1074/jbc.M110.102699Search in Google Scholar

Loschen, G., Flohe, L., and Chance, B. (1971). Respiratory chain linked H2O2 production in pigeon heart mitochondria. FEBS Lett. 18, 261–264.10.1016/0014-5793(71)80459-3Search in Google Scholar

Lu, J. and Holmgren, A. (2014). The thioredoxin antioxidant system. Free Radic. Biol. Med. 66, 75–87.10.1016/j.freeradbiomed.2013.07.036Search in Google Scholar PubMed

Macian, F. (2005). NFAT proteins: key regulators of T-cell development and function. Nat. Rev. Immunol. 5, 472–484.10.1038/nri1632Search in Google Scholar PubMed

Mailloux, R.J. (2015). Teaching the fundamentals of electron transfer reactions in mitochondria and the production and detection of reactive oxygen species. Redox Biol. 4, 381–398.10.1016/j.redox.2015.02.001Search in Google Scholar PubMed PubMed Central

Mailloux, R.J. (2016). Application of mitochondria-targeted pharmaceuticals for the treatment of heart disease. Curr. Pharm. Des. 22, 4763–4779.10.2174/1381612822666160629070914Search in Google Scholar PubMed

Mailloux, R.J. and Harper, M.E. (2011). Uncoupling proteins and the control of mitochondrial reactive oxygen species production. Free Radic. Biol. Med. 51, 1106–1115.10.1016/j.freeradbiomed.2011.06.022Search in Google Scholar PubMed

Mailloux, R.J. and Willmore, W.G. (2014). S-glutathionylation reactions in mitochondrial function and disease. Front. Cell Dev. Biol. 2, 68.10.3389/fcell.2014.00068Search in Google Scholar PubMed PubMed Central

Mailloux, R.J. and Treberg, J.R. (2016). Protein S-glutathionlyation links energy metabolism to redox signaling in mitochondria. Redox Biol. 8, 110–118.10.1016/j.redox.2015.12.010Search in Google Scholar PubMed PubMed Central

Mailloux, R.J., Seifert, E.L., Bouillaud, F., Aguer, C., Collins, S., and Harper, M.E. (2011). Glutathionylation acts as a control switch for uncoupling proteins UCP2 and UCP3. J. Biol. Chem. 286, 21865–21875.10.1074/jbc.M111.240242Search in Google Scholar PubMed PubMed Central

Mailloux, R.J., Adjeitey, C.N., Xuan, J.Y., and Harper, M.E. (2012a). Crucial yet divergent roles of mitochondrial redox state in skeletal muscle vs. brown adipose tissue energetics. FASEB J. 26, 363–375.10.1096/fj.11-189639Search in Google Scholar PubMed

Mailloux, R.J., Fu, A., Robson-Doucette, C., Allister, E.M., Wheeler, M.B., Screaton, R., and Harper, M.E. (2012b). Glutathionylation state of uncoupling protein-2 and the control of glucose-stimulated insulin secretion. J. Biol. Chem. 287, 39673–39685.10.1074/jbc.M112.393538Search in Google Scholar PubMed PubMed Central

Mailloux, R.J., Xuan, J.Y., Beauchamp, B., Jui, L., Lou, M., and Harper, M.E. (2013). Glutaredoxin-2 is required to control proton leak through uncoupling protein-3. J. Biol. Chem. 288, 8365–8379.10.1074/jbc.M112.442905Search in Google Scholar PubMed PubMed Central

Mailloux, R.J., Xuan, J.Y., McBride, S., Maharsy, W., Thorn, S., Holterman, C.E., Kennedy, C.R., Rippstein, P., deKemp, R., da Silva, J., et al. (2014). Glutaredoxin-2 is required to control oxidative phosphorylation in cardiac muscle by mediating deglutathionylation reactions. J. Biol. Chem. 289, 14812–14828.10.1074/jbc.M114.550574Search in Google Scholar PubMed PubMed Central

Mailloux, R.J., Craig Ayre, D., and Christian, S.L. (2016a). Induction of mitochondrial reactive oxygen species production by GSH mediated S-glutathionylation of 2-oxoglutarate dehydrogenase. Redox Biol. 8, 285–297.10.1016/j.redox.2016.02.002Search in Google Scholar PubMed PubMed Central

Mailloux, R.J., Gardiner, D., and O’Brien, M. (2016b). 2-Oxoglutarate dehydrogenase is a more significant source of O2˙/H2O2 than pyruvate dehydrogenase in cardiac and liver tissue. Free Radic. Biol. Med. 97, 501–512.10.1016/j.freeradbiomed.2016.06.014Search in Google Scholar PubMed

Mannervik, B. and Axelsson, K. (1980). Role of cytoplasmic thioltransferase in cellular regulation by thiol-disulphide interchange. Biochem. J. 190, 125–130.10.1042/bj1900125Search in Google Scholar PubMed PubMed Central

Maranzana, E., Barbero, G., Falasca, A.I., Lenaz, G., and Genova, M.L. (2013). Mitochondrial respiratory supercomplex association limits production of reactive oxygen species from complex I. Antioxid. Redox Signal. 19, 1469–1480.10.1089/ars.2012.4845Search in Google Scholar PubMed PubMed Central

McLain, A.L., Szweda, P.A., and Szweda, L.I. (2011). α-Ketoglutarate dehydrogenase: a mitochondrial redox sensor. Free Radic. Res. 45, 29–36.10.3109/10715762.2010.534163Search in Google Scholar PubMed PubMed Central

McLain, A.L., Cormier, P.J., Kinter, M., and Szweda, L.I. (2013). Glutathionylation of α-ketoglutarate dehydrogenase: the chemical nature and relative susceptibility of the cofactor lipoic acid to modification. Free Radic. Biol. Med. 61, 161–169.10.1016/j.freeradbiomed.2013.03.020Search in Google Scholar PubMed PubMed Central

Meier, B., Radeke, H.H., Selle, S., Younes, M., Sies, H., Resch, K., and Habermehl, G.G. (1989). Human fibroblasts release reactive oxygen species in response to interleukin-1 or tumour necrosis factor-α. Biochem. J. 263, 539–545.10.1042/bj2630539Search in Google Scholar PubMed PubMed Central

Moreno-Loshuertos, R. and Enriquez, J.A. (2016). Respiratory supercomplexes and the functional segmentation of the CoQ pool. Free Radic. Biol. Med. 100, 5–13.10.1016/j.freeradbiomed.2016.04.018Search in Google Scholar PubMed

Murphy, M.P. (2009). How mitochondria produce reactive oxygen species. Biochem. J. 417, 1–13.10.1042/BJ20081386Search in Google Scholar PubMed PubMed Central

Murphy, M.P. (2012). Mitochondrial thiols in antioxidant protection and redox signaling: distinct roles for glutathionylation and other thiol modifications. Antioxid. Redox Signal. 16, 476–495.10.1089/ars.2011.4289Search in Google Scholar PubMed

Nagy, P., Wang, X., Lemma, K., and Ashby, M.T. (2007). Reactive sulfur species: hydrolysis of hypothiocyanite to give thiocarbamate-S-oxide. J. Am. Chem. Soc. 129, 15756–15757.10.1021/ja0770532Search in Google Scholar PubMed

Nakamura, Y., Colburn, N.H., and Gindhart, T.D. (1985). Role of reactive oxygen in tumor promotion: implication of superoxide anion in promotion of neoplastic transformation in JB-6 cells by TPA. Carcinogenesis 6, 229–235.10.1093/carcin/6.2.229Search in Google Scholar PubMed

Negre-Salvayre, A., Hirtz, C., Carrera, G., Cazenave, R., Troly, M., Salvayre, R., Penicaud, L., and Casteilla, L. (1997). A role for uncoupling protein-2 as a regulator of mitochondrial hydrogen peroxide generation. FASEB J. 11, 809–815.10.1096/fasebj.11.10.9271366Search in Google Scholar

Nemeria, N.S., Ambrus, A., Patel, H., Gerfen, G., Adam-Vizi, V., Tretter, L., Zhou, J., Wang, J., and Jordan, F. (2014). Human 2-oxoglutarate dehydrogenase complex E1 component forms a thiamin-derived radical by aerobic oxidation of the enamine intermediate. J. Biol. Chem. 289, 29859–29873.10.1074/jbc.M114.591073Search in Google Scholar

Nulton-Persson, A.C., Starke, D.W., Mieyal, J.J., and Szweda, L.I. (2003). Reversible inactivation of α-ketoglutarate dehydrogenase in response to alterations in the mitochondrial glutathione status. Biochemistry 42, 4235–4242.10.1021/bi027370fSearch in Google Scholar

O’Brien, M., Chalker, J., Slade, L., Gardiner, D., and Mailloux, R.J. (2017). Protein S-glutathionylation alters superoxide/hydrogen peroxide emission from pyruvate dehydrogenase complex. Free Radic. Biol. Med. 106, 302–314.10.1016/j.freeradbiomed.2017.02.046Search in Google Scholar

Pai, H.V., Starke, D.W., Lesnefsky, E.J., Hoppel, C.L., and Mieyal, J.J. (2007). What is the functional significance of the unique location of glutaredoxin 1 (GRx1) in the intermembrane space of mitochondria? Antioxid. Redox Signal. 9, 2027–2033.10.1089/ars.2007.1642Search in Google Scholar

Papaconstantinou, J. (2009). Insulin/IGF-1 and ROS signaling pathway cross-talk in aging and longevity determination. Mol. Cell. Endocrinol. 299, 89–100.10.1016/j.mce.2008.11.025Search in Google Scholar

Perkins, A., Poole, L.B., and Karplus, P.A. (2014). Tuning of peroxiredoxin catalysis for various physiological roles. Biochemistry 53, 7693–7705.10.1021/bi5013222Search in Google Scholar

Quinlan, C.L., Orr, A.L., Perevoshchikova, I.V., Treberg, J.R., Ackrell, B.A., and Brand, M.D. (2012). Mitochondrial complex II can generate reactive oxygen species at high rates in both the forward and reverse reactions. J. Biol. Chem. 287, 27255–27264.10.1074/jbc.M112.374629Search in Google Scholar

Quinlan, C.L., Goncalves, R.L., Hey-Mogensen, M., Yadava, N., Bunik, V.I., and Brand, M.D. (2014). The 2-oxoacid dehydrogenase complexes in mitochondria can produce superoxide/hydrogen peroxide at much higher rates than complex I. J. Biol. Chem. 289, 8312–8325.10.1074/jbc.M113.545301Search in Google Scholar

Radi, R., Turrens, J.F., Chang, L.Y., Bush, K.M., Crapo, J.D., and Freeman, B.A. (1991). Detection of catalase in rat heart mitochondria. J. Biol. Chem. 266, 22028–22034.10.1016/S0021-9258(18)54740-2Search in Google Scholar

Ramsden, D.B., Ho, P.W., Ho, J.W., Liu, H.F., So, D.H., Tse, H.M., Chan, K.H., and Ho, S.L. (2012). Human neuronal uncoupling proteins 4 and 5 (UCP4 and UCP5): structural properties, regulation, and physiological role in protection against oxidative stress and mitochondrial dysfunction. Brain Behav. 2, 468–478.10.1002/brb3.55Search in Google Scholar PubMed PubMed Central

Rhee, S.G. (2006). Cell signaling. H2O2, a necessary evil for cell signaling. Science 312, 1882–1883.10.1126/science.1130481Search in Google Scholar PubMed

Rhee, S.G. and Kil, I.S. (2016). Mitochondrial H2O2 signaling is controlled by the concerted action of peroxiredoxin III and sulfiredoxin: linking mitochondrial function to circadian rhythm. Free Radic. Biol. Med. 99, 120–127.10.1016/j.freeradbiomed.2016.10.011Search in Google Scholar PubMed

Rhee, S.G. and Kil, I.S. (2017). Multiple Functions and regulation of mammalian peroxiredoxins. Annu. Rev. Biochem. 86, 749–775.10.1146/annurev-biochem-060815-014431Search in Google Scholar PubMed

Ricoult, S.J. and Manning, B.D. (2013). The multifaceted role of mTORC1 in the control of lipid metabolism. EMBO Rep. 14, 242–251.10.1038/embor.2013.5Search in Google Scholar PubMed PubMed Central

Robson-Doucette, C.A., Sultan, S., Allister, E.M., Wikstrom, J.D., Koshkin, V., Bhattacharjee, A., Prentice, K.J., Sereda, S.B., Shirihai, O.S., and Wheeler, M.B. (2011). Beta-cell uncoupling protein 2 regulates reactive oxygen species production, which influences both insulin and glucagon secretion. Diabetes 60, 2710–2719.10.2337/db11-0132Search in Google Scholar PubMed PubMed Central

Rupprecht, A., Brauer, A.U., Smorodchenko, A., Goyn, J., Hilse, K.E., Shabalina, I.G., Infante-Duarte, C. and Pohl, E.E. (2012). Quantification of uncoupling protein 2 reveals its main expression in immune cells and selective up-regulation during T-cell proliferation. PLoS One 7, e41406.10.1371/journal.pone.0041406Search in Google Scholar PubMed PubMed Central

Sadidi, M., Lentz, S.I., and Feldman, E.L. (2009). Hydrogen peroxide-induced Akt phosphorylation regulates Bax activation. Biochimie 91, 577–585.10.1016/j.biochi.2009.01.010Search in Google Scholar PubMed PubMed Central

Schagger, H. and Pfeiffer, K. (2000). Supercomplexes in the respiratory chains of yeast and mammalian mitochondria. EMBO J. 19, 1777–1783.10.1093/emboj/19.8.1777Search in Google Scholar PubMed PubMed Central

Schrauwen, P., Hoeks, J., Schaart, G., Kornips, E., Binas, B., Van De Vusse, G.J., Van Bilsen, M., Luiken, J.J., Coort, S.L., Glatz, J.F., et al. (2003). Uncoupling protein 3 as a mitochondrial fatty acid anion exporter. FASEB J. 17, 2272–2274.10.1096/fj.03-0515fjeSearch in Google Scholar PubMed

Sena, L.A., Li, S., Jairaman, A., Prakriya, M., Ezponda, T., Hildeman, D.A., Wang, C.R., Schumacker, P.T., Licht, J.D., Perlman, H., et al. (2013). Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity 38, 225–236.10.1016/j.immuni.2012.10.020Search in Google Scholar PubMed PubMed Central

Shabalina, I.G., Vrbacky, M., Pecinova, A., Kalinovich, A.V., Drahota, Z., Houstek, J., Mracek, T., Cannon, B., and Nedergaard, J. (2014). ROS production in brown adipose tissue mitochondria: the question of UCP1-dependence. Biochim. Biophys. Acta 1837, 2017–2030.10.1016/j.bbabio.2014.04.005Search in Google Scholar PubMed

Shelton, M.D., Chock, P.B., and Mieyal, J.J. (2005). Glutaredoxin: role in reversible protein s-glutathionylation and regulation of redox signal transduction and protein translocation. Antioxid. Redox Signal. 7, 348–366.10.1089/ars.2005.7.348Search in Google Scholar PubMed

Sies, H. (1985). Oxidative stress: introductory remarks. In: Oxidative Stress, H. Sies, ed. (Academic Press, London), pp. 1–8.10.1016/B978-0-12-642760-8.50005-3Search in Google Scholar

Sies, H., Berndt, C., and Jones, D.P. (2017). Oxidative stress. Annu. Rev. Biochem. 86, 715–748.10.1146/annurev-biochem-061516-045037Search in Google Scholar PubMed

Slade, L., Chalker, J., Kuksal, N., Young, A., Gardiner, D., and Mailloux, R.J. (2017). Examination of the superoxide/hydrogen peroxide forming and quenching potential of mouse liver mitochondria. Biochim. Biophys. Acta. 1861, 1960–1969.10.1016/j.bbagen.2017.05.010Search in Google Scholar PubMed

Sobotta, M.C., Liou, W., Stocker, S., Talwar, D., Oehler, M., Ruppert, T., Scharf, A.N., and Dick, T.P. (2015). Peroxiredoxin-2 and STAT3 form a redox relay for H2O2 signaling. Nat. Chem. Biol. 11, 64–70.10.1038/nchembio.1695Search in Google Scholar PubMed

Taylor, E.R., Hurrell, F., Shannon, R.J., Lin, T.K., Hirst, J., and Murphy, M.P. (2003). Reversible glutathionylation of complex I increases mitochondrial superoxide formation. J. Biol. Chem. 278, 19603–19610.10.1074/jbc.M209359200Search in Google Scholar PubMed

Tormos, K.V., Anso, E., Hamanaka, R.B., Eisenbart, J., Joseph, J., Kalyanaraman, B., and Chandel, N.S. (2011). Mitochondrial complex III ROS regulate adipocyte differentiation. Cell Metab. 14, 537–544.10.1016/j.cmet.2011.08.007Search in Google Scholar PubMed PubMed Central

Trenker, M., Malli, R., Fertschai, I., Levak-Frank, S., and Graier, W.F. (2007). Uncoupling proteins 2 and 3 are fundamental for mitochondrial Ca2+ uniport. Nat. Cell Biol. 9, 445–452.10.1038/ncb1556Search in Google Scholar PubMed PubMed Central

Tretter, L. and Adam-Vizi, V. (1999). Inhibition of α-ketoglutarate dehydrogenase due to H2O2-induced oxidative stress in nerve terminals. Ann. N. Y. Acad. Sci. 893, 412–416.10.1111/j.1749-6632.1999.tb07867.xSearch in Google Scholar PubMed

Tretter, L. and Adam-Vizi, V. (2004). Generation of reactive oxygen species in the reaction catalyzed by α-ketoglutarate dehydrogenase. J Neurosci 24, 7771–7778.10.1523/JNEUROSCI.1842-04.2004Search in Google Scholar PubMed PubMed Central

Tretter, L. and Adam-Vizi, V. (2005). α-Ketoglutarate dehydrogenase: a target and generator of oxidative stress. Philos. Trans. R. Soc. Lond. B Biol. Sci. 360, 2335–2345.10.1098/rstb.2005.1764Search in Google Scholar PubMed PubMed Central

Verkhovskaya, M. and Bloch, D.A. (2013). Energy-converting respiratory Complex I: on the way to the molecular mechanism of the proton pump. Int. J. Biochem. Cell. Biol. 45, 491–511.10.1016/j.biocel.2012.08.024Search in Google Scholar PubMed

Weinberg, S.E., Sena, L.A., and Chandel, N.S. (2015). Mitochondria in the regulation of innate and adaptive immunity. Immunity 42, 406–417.10.1016/j.immuni.2015.02.002Search in Google Scholar PubMed PubMed Central

Wilson-Fritch, L., Burkart, A., Bell, G., Mendelson, K., Leszyk, J., Nicoloro, S., Czech, M., and Corvera, S. (2003). Mitochondrial biogenesis and remodeling during adipogenesis and in response to the insulin sensitizer rosiglitazone. Mol. Cell. Biol. 23, 1085–1094.10.1128/MCB.23.3.1085-1094.2003Search in Google Scholar PubMed PubMed Central

Yoo, S.E., Chen, L., Na, R., Liu, Y., Rios, C., Van Remmen, H., Richardson, A., and Ran, Q. (2012). Gpx4 ablation in adult mice results in a lethal phenotype accompanied by neuronal loss in brain. Free Radic. Biol. Med. 52, 1820–1827.10.1016/j.freeradbiomed.2012.02.043Search in Google Scholar PubMed PubMed Central

Yun, J. and Finkel, T. (2014). Mitohormesis. Cell Metab. 19, 757–766.10.1016/j.cmet.2014.01.011Search in Google Scholar PubMed PubMed Central

Zhang, J., Khvorostov, I., Hong, J.S., Oktay, Y., Vergnes, L., Nuebel, E., Wahjudi, P.N., Setoguchi, K., Wang, G., Do, A., et al. (2011). UCP2 regulates energy metabolism and differentiation potential of human pluripotent stem cells. EMBO J. 30, 4860–4873.10.1038/emboj.2011.401Search in Google Scholar PubMed PubMed Central

Zhu, H., Santo, A., and Li, Y. (2012). The antioxidant enzyme peroxiredoxin and its protective role in neurological disorders. Exp. Biol. Med. (Maywood) 237, 143–149.10.1258/ebm.2011.011152Search in Google Scholar PubMed

Received: 2017-5-10
Accepted: 2017-6-27
Published Online: 2017-7-4
Published in Print: 2017-10-26

©2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 23.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/hsz-2017-0160/html
Scroll to top button