Abstract
Mitochondria are the main intracellular location for fuel generation; however, they are not just power plants but involved in a range of other intracellular functions including regulation of redox homeostasis and cell fate. Dysfunction of mitochondria will result in oxidative stress which is one of the underlying causal factors for a variety of diseases including neurodegenerative diseases, diabetes, cardiovascular diseases, and cancer. In this paper, generation of reactive oxygen/nitrogen species (ROS/RNS) in the mitochondria, redox regulatory roles of certain mitochondrial proteins, and the impact on cell fate will be discussed. The current state of our understanding in mitochondrial dysfunction in pathological states and how we could target them for therapeutic purpose will also be briefly reviewed.
1. Introduction
Mitochondria are the main intracellular location for generating adenosine triphosphate (ATP), the fuel for cell's metabolic needs, and therefore are referred to as the power plants of the cell. Energy is stored in the form of phosphate bond and is released when ATP is hydrolyzed to adenosine diphosphate (ADP) to meet the requirement of a number of energy demanding cellular processes. ATP is generated through cellular respiration, including a set of chemical reactions named as the tricarboxylic acid (TCA) cycle and oxidative phosphorylation that take place in mitochondria; therefore, mitochondria play a crucial regulatory role in cellular metabolism [1]. However, mitochondria are far more than just power suppliers. They are also involved in many other cellular functions, including calcium signaling [2], heme [3] and steroid synthesis [4], regulation of membrane potential [1], proliferation [5] or apoptosis [6], and redox homeostasis maintenance [7], to name just a few. Mitochondria are the major sites for free radical species production, including both reactive oxygen species (ROS) and reactive nitrogen species (RNS). On one hand, free radical species are indispensable for proper cell signaling; on the other hand, excessive generation of ROS results in cell/tissue injury and death. In this paper, the underlying mechanisms for generation of free radical species in the mitochondria and how some mitochondrial proteins act as redox regulators will be of prime focus. In the past decade, more and more pieces of evidence are surfacing to point the role of ROS as critical mediators of the balance between cell proliferation and cell death [8–10]. Therefore, the involvement of mitochondria in cell death, especially from the noncanonical view where mitochondria regulate cell death through manipulating the redox millieu will also be reviewed. Due to the fundamental regulatory role of mitochondria, disturbances of the integrity of their functions will result in cellular dysfunction leading to various pathological states or even death. Therefore, a good understanding of the basic mitochondrial biology will help in therapeutic design for better disease management.
2. Mitochondrial Functions
2.1. Mitochondria Structure and Metabolism
2.1.1. Mitochondria Structure
Mitochondria are essential to sustain life as around 98% of the oxygen that we breathe in is consumed by mitochondria for energy production. In order to understand the basic principles of mitochondrial bioenergetics, it is necessary to have a brief overview of its structure first. Mitochondria, as thought to have evolved from a bacterial progenitor and having their own mitochondrial DNA pool [11], are bounded by two membrane systems, including an outer and an inner membrane, and the space in between is referred to as the intermembrane space; however, the two layers of membranes occasionally come into contact with each other to form junctional complexes. The inner mitochondrial membrane, with multiple inward folding known as cristae which house membrane-bound mitochondrial enzymes, serves as the major barrier between mitochondria and cytoplasm since it is largely impermeable, thereby preventing small molecules and ionic species from entering the mitochondrial matrix [1].
2.1.2. TCA Cycle and Electron Transport Chain
The tricarboxylic acid (TCA) cycle and the electron transport chain contribute to the key enzymatic components of the mitochondria. During the process of breaking down carbon substrates into acetyl CoA, reducing equivalents (NADH and FADH) are produced, which are then fed to the electron transport chain consisting of Complex I (NADH dehydrogenase), Complex II (succinate dehydrogenase), Complex III (ubiquinol cytochrome c reductase), and Complex IV (cytochrome c oxidase). Electrons move from the reducing equivalents to Complexes I and II, respectively, which are then passed onto ubisemiquinone for shuttling to Complex III followed by Complex IV through cytochrome c. In the meanwhile, an electrochemical proton gradient is generated when protons are transferred across the inner mitochondrial membrane into the intermembrane space coupled to the electron transfer at Complexes I, III, and IV. This is known as the proton-motive force, generating the mitochondrial transmembrane potential, which is usually about negative 150–180 mV as compared to the cytoplasm. The influx of the protons through the proton translocating F1F0-ATP synthase, driven by this force, is coupled to a chemical reaction that phosphorylates ADP to ATP [12]. An electrogenic transporter, adenine nucleotide translocase (ANT), then transports ATP out of the mitochondria to places where energy demanding cellular processes occur [1].
2.2. Mitochondria and Redox Homeostasis Maintenance
2.2.1. Sources of Mitochondrial Reactive Radical Species
Reactive Oxygen Species (ROS) —
Mitochondria are the major sources of free radical species production since unpaired electrons are generated in the process of oxidative phosphorylation. Partial reduction of molecular oxygen by the unpaired electrons leads to the production of superoxide anions (O2 ∙−) which are one of the reactive oxygen species (ROS) and readily converted to hydrogen peroxide (H2O2) by magnesium superoxide dismutase (MnSOD) residing in mitochondria matrix. H2O2 can be converted subsequently to the highly reactive oxygen species, hydroxyl radical (∙OH), through the Fenton reaction [7]. Unlike H2O2, O2 ∙− does not diffuse that readily across the membrane and thus for O2 ∙− produced in the matrix, the activity of MnSOD is critical to prevent the mitochondrial matrix components from oxidative damage.
Although, it is well known that ROS can be generated as byproducts of oxidative phosphorylation, the question as to the specific site(s) along the electron transport chain responsible for ROS generation has always been a hotly debated topic. It is traditionally believed that under physiological conditions, Complex I is the main site for mitochondrial ROS production, where O2 ∙− is produced on the matrix side and rapidly dismutated to H2O2 [13, 14]. In addition, Complex III has also been reported as a site for O2 ∙− production [15, 16]. It is demonstrated that under ischemic and apoptotic conditions, O2 ∙− production is triggered at Complex III. This may happen through inhibition of Complex IV as well as over reduction of the electron transport chain in the event of mounting hypoxic stress [17]. In a more recent review [18], the relative contribution of each complex towards O2 ∙− production has been clearly quantitated with Complexes I (producing O2 ∙− to the matrix) and III (producing O2 ∙− to both the matrix and intermembrane space) having the greatest maximum capacities while Complex II has normally negligible rates.
Another major source of ROS production in the cell is the NADPH oxidase (Nox) family proteins, which are enzyme complexes catalyzing the electron transfer from NADPH to molecular oxygen, generating O2 ∙− and H2O2. The relative quantitative contribution of mitochondria and NADPH oxidase in cellular ROS production is expected to vary greatly from one cell type to another. In certain cells including the phagocytic neutrophils as well as nonphagocytic fibroblasts, vascular smooth muscle cells and endothelial cells, cellular ROS production is largely contributed by NADPH oxidase [18–20]. It has been recently reported that one of the Nox isoforms, Nox4, is expressed in the mitochondria in the rat renal cortex [21], cardiac myocytes [22, 23], and in the mitochondria-enriched heavy membrane fractions of the chronic myeloid leukemia cells that overexpress the antiapoptotic protein Bcl-2 (CEM/Bcl-2) [24]. Nox4 expression was shown to correlate with dihydroethidium staining for O2 ∙−. In Nox4-transgenic mice cysteine residues of mitochondrial proteins were also more oxidized [23]. However, there are no direct and specific measurements of Nox4 activity in the mitochondria up to date. Meanwhile, the cytoplasmic Nox4 may be involved in the activation of PKCε, mitoKATP and modulation of thioredoxin 2 activity resulting in redox-sensitive upregulation of mitochondrial ROS production through the electron transport chain [25]. Apart from the electron transport chain and the NOX family, several other sites in the mitochondria have also been reported to generate O2 ∙−, including pyruvate dehydrogenase, α-ketaglutarate dehydrogenase [26], glycerol-3-phosphate dehydrogenase, and fatty acid β-oxidation [18].
Recently, new findings that supported the concept of mitochondrial O2 ∙− flashes have gained much interest and revealed several previously unknown aspects of mitochondrial dynamics. Transient quantal O2 ∙− flashes were observed in excitable cells such as muscle cells and neurons in vivo and they are associated with mitochondrial permeability transition pore (mPTP) opening, which presents a new facet in physiological ROS production [27–29].
Reactive Nitrogen Species (RNS) —
Apart from ROS generation, there is evidence to suggest the expression of a mitochondrial specific nitric oxide synthase (NOS) leading to ∙NO production. It was first described in 1997 by Ghafourifar and Richter that ∙NO was generated from isolated mitochondria when they were loaded with calcium and mitochondrial potential dropped in some of the treated mitochondria [30]. Another group provided more direct evidence by showing images of calcium regulated ∙NO production within the mitochondria of permeabilized cells, through the use of the ∙NO-sensitive chromophore, DAF-2 [31].
The impact of ∙NO generation on mitochondrial functions remains controversial, largely depending on the amount of ∙NO that is produced as well as the conditions under which it is produced. It has been shown that mitochondrial respiration can be partially inhibited with even the modest levels of ∙NO causing an increase in mitochondrial ROS production. This also results in mitochondrial depolarization resulting in a decrease in mitochondrial calcium uptake [32]. The inhibitory effect of ∙NO on mitochondrial respiration has been shown to mainly result from inactivation of cytochrome c oxidase (Complex IV) [33–40], which is the rate-limiting step of the electron transport chain. Being a double-edged sword, ∙NO was demonstrated to play a role in mitochondrial biogenesis as well, which gets stimulated by ∙NO generation through a cGMP-dependent upregulation of PGC1α expression, which in turn increases the expression of mtTFA and NRF-1 resulting in an increased mitochondria biosynthesis observed in adipocytes and hepatocytes [41]. Another group also showed a physiological role for ∙NO generation in mitochondrial mass regulation where endothelial-∙NO-synthase- (eNOS-) deficient mice showed deficiencies in mitochondrial enzymes [42]. In addition, protective effect of ∙NO on mitochondria and cells is also observed against ischemia/reperfusion (I/R) injury (late preconditioning). It has been suggested that partial temporary inhibition of Complex I might be involved in this effect via inhibition of ROS burst in reperfusion. Furthermore, treatment with diet inorganic nitrates or administration of nitrites or S-nitroso-2-mercaptopropionyl glycine (SNO-MPG) confers cardioprotection against the injury via temporary S-nitrosylation of various cellular protein targets [43].
On the contrary, under some pathological stimuli, excessive production of ∙NO could result in severe tissue damage. Several groups have suggested that in the face of a high concentration of ∙NO with a relatively low oxygen concentration, respiration can be inhibited resulting in an increased O2 ∙− production, which will react with the freely membrane diffusible ∙NO to form a much more reactive radical species, peroxynitrite (ONOO−). Peroxynitrite can also be formed through an alternative route via the reaction of nitroxyl anion (NO−) and molecular oxygen [44]. Nitroxyl anion can be derived from ∙NO through one electron reduction by the electron donors including cytochrome c [45] or ubiquinol [46]. Due to its ability to diffuse across the mitochondrial membranes, ONOO− can result in oxidative damage of critical components throughout the mitochondria via oxidation, nitration, and/or nitrosation. For example, thiols present in Complex I can be oxidized leading to the formation of an S-nitrosothiol derivative and inactivation of the complex [47–49]. Complexes II and V were also shown to be inactivated by ONOO− [34, 44, 48–53]. MnSOD, the matrix antioxidant enzyme, is also a target for nitration leading to a decrease in its enzymatic activity [54]. Adenine nucleotide translocase (ANT) [55], creatine kinase [56], nicotinamide nucleotide transhydrogenase [57], aconitase [58], and components of the pyridine nucleotide-dependent calcium release pathway [59] are all targets of ONOO−. Therefore, ONOO− has profound effects on mitochondrial metabolism, calcium homeostasis, and the mitochondrial permeability transition pore [42]. In addition, ONOO− has been shown to uncouple eNOS, leading to a switch from ∙NO to O2 ∙− production and an increase in mitochondrial ROS levels as well [42, 60].
Of particular note, carbon monoxide (CO), an endogeneous gas produced by heme oxygenase (HO) that catalyzes heme degradation, is able to induce the production of ROS and RNS as well due to its high affinity for reduced transition metals such as Fe2+. For example, it binds to Complex IV and slows the terminal transfer rate of electrons to molecular O2 leading to enhanced O2 ∙− production [61].
2.2.2. Mitochondria in Redox Regulation
Since mitochondria are major sources for ROS production, it is not surprising that they are well equipped with antioxidant defenses, including a large pool of glutathione, glutathione peroxidase, glutathione reductase, MnSOD, catalase, and the thioredoxin system [1, 62].
Although excessive levels of ROS will lead to protein oxidation and lipid peroxidation causing damage to mitochondrial membrane, proteins, and DNA, especially when the mitochondrial DNA is not protected with associated histones, lower levels of ROS have been demonstrated to be essential signaling molecules [7, 63, 64]. A new concept is now emerging that mitochondrial ROS production is likely to be highly regulated as a part of physiological mitochondrial functions and the underlying molecular mechanisms are being gradually uncovered [7]. In this paper, a few mitochondrial proteins that act as redox regulators will be discussed as examples, including the antiapoptotic protein Bcl-2, cytochrome c oxidase (COX), and the small GTPase Rac1.
Bcl-2 and Its Effect on Mitochondrial ROS Generation —
Bcl-2, one of the antiapoptotic members of the Bcl-2 family proteins residing on the outer mitochondrial membrane [65], is best known for its ability to inhibit apoptotic execution, that is, to form homo- and heterodimers to prevent the oligomerization of proapoptotic Bcl-2 family members, thus antagonizing the induction of mitochondrial outer membrane permeabilization (MOMP) [66]. However, recent evidence points to a new facet of Bcl-2 biology in redox regulation. The involvement of Bcl-2 in redox regulation was first demonstrated by Hockenbery et al. that Bcl-2 overexpression protected against ROS-induced apoptosis [67]. Soon after other studies also revealed the protective capacity of Bcl-2 against various ROS triggers [68–73]. However, Bcl-2 itself was later found out to possess no intrinsic antioxidant ability [74] implying that rather its overexpression indirectly induces an enhancement in antioxidant capacity of the cells when they undergo overt oxidative stress [66, 75] and this is supported by the findings that Bcl-2-mediated protection is associated with upregulation of the cellular enzymatic and nonenzymatic antioxidant defense machineries including the glutathione system, catalase, and NAD(P)H [74, 76–79]. More recently, a clearer picture of Bcl-2 with respect to its ability to regulate redox status is emerging, and a prooxidant role of Bcl-2 is established under normal physiological states [77, 80–84]. Based on this prooxidant property of Bcl-2, it implies that the enhanced antioxidant capacity that was observed with Bcl-2 overexpression could be an adaptive response to the chronic but mild oxidative intracellular milieu [70, 74, 76, 79, 83] and this serves as a first-line defense in the event of acute oxidative insults maintaining the ROS levels within a threshold optimal for cell survival [66, 75, 81, 85].
The underlying mechanisms on how Bcl-2 exerts its prooxidant activity, however, have not been fully elucidated. It was first hypothesized that the prooxidant milieu in Bcl-2 overexpressing mitochondria resulted from an altered dynamics of the oxidative phosphorylation. An increase in mitochondrial size and associated matrix content was observed with Bcl-2 overexpression and this indicated an increase in the number of electron donors and a subsequent increase in the chance of electrons leaking out of the electron transport chain to form O2 ∙− [77, 86]. However, the exact mechanism linking Bcl-2 expression levels to mitochondrial size and matrix content was not addressed in the above-mentioned studies. More recently, our group has established the inherent ability of Bcl-2 to generate intramitochondrial O2 ∙− by engaging mitochondrial respiration in tumor cells. An increased mitochondrial oxygen consumption rate and cytochrome c oxidase (COX or Complex IV) activity was observed in Bcl-2 overexpressing cells [81, 85, 87]. It is plausible that the increased mitochondrial respiration rate results in an increased electron flux across the electron transport chain and an increased probability of leakage of electrons onto molecular oxygen thus leading to an increase in the by-production of O2 ∙−. Indeed, either silencing of Bcl-2 with siRNA or functional inhibition of Bcl-2 with the BH3 mimetic, HA14-1, in those cells, reversed both the oxygen consumption rate as well as the O2 ∙− levels [81]. This is further supported by the observation that mitochondrial respiratory rate and O2 ∙− levels correlated with Bcl-2 expression levels across different tumor cell lines with various endogeneous Bcl-2 levels [85]. Of note, Bcl-2 overexpression promoted the mitochondrial localization of COX Va and Vb, which are nuclear encoded subunits of Complex IV, which could explain for the significantly increased Complex IV activity in these cells [85]; it has been previously shown that mitochondrial level of COX Vb correlated with the COX holoenzyme activity [88]. Of particular note, the increased O2 ∙− release as induced by Bcl-2 overexpression might seem contradictory to what we mentioned in the earlier section that inhibition of ETC by ∙NO also results in O2 ∙− release. However, the former occurs as a result of increased electron flux across the electron transport chain and an increased probability of leakage of electrons onto molecular oxygen while the latter is the result of inhibition of reduction step along the ETC leading to promotion of the reaction of oxygen with accumulated reductants [89].
More recently, our group has identified another functional player in Bcl-2-mediated prooxidant state, the small GTPase Rac1 [24]. Rac1 is known to be involved in the assembly and activation of NADPH oxidase complex leading to O2 ∙− production. It was first discovered that introduction of the dominant negative mutant Rac1N17 neutralized the prooxidant activity of Bcl-2 [82]. Later on, a physical interaction was observed between these two proteins in the outer mitochondrial membrane of tumor cells which could be blocked with the BH3 mimetics and Bcl-2 BH3 domain peptides. The intramitochondrial O2 ∙− production in Bcl-2 overexpressing cells was also reversed by BH3 peptides, which can also be achieved with silencing or functional inhibition of Rac1 [24].
These data provide evidence for the existence of functional complexes within the mitochondria involving Bcl-2, but the precise mechanism of interaction and how disruption of these interaction(s) could impact cell fate remains to be elucidated. Of course, apart from the above-mentioned proteins, there are many others that act as mitochondrial redox regulators including the master transcription factor for cellular antioxidant defense machinery, the transcription factor NF-E2-related factor 2 (Nrf2), which gets activated under cellular oxidative stress conditions such as GSH depletion, ∙NO, and nitrosative stress [90–92]. Interestingly, a recent paper provided the first evidence that activation of Nrf2 can upregulate Bcl-2 as well [93].
2.3. Mitochondria and Cell Fate Regulation
Since mitochondria are fundamental energy generators, severe damage to mitochondria will inevitably cause disorders in cellular functions. Once ionic gradients and intracellular osmolarity cannot be maintained, cells will swell and die through a death process known as necrosis [94]. However, for some immortalized cells, they survive reasonably well on ATP generated from glycolysis even when mitochondrial respiration is completely inhibited [1, 95, 96]. Apart from dying passively when the ATP supply fails, cells can also actively undergo a “suicide” program through the mitochondria-mediated apoptotic pathway upon compromise in the mitochondrial outer and/or inner membrane permeability [97]. Furthermore, there is another form of cell death named autophagy, which degrades cellular organelles and proteins promoting either survival or death depending on the stress conditions. Various studies have indicated the involvement of ROS and mitochondria in autophagic regulation [98, 99]; however, due to space constraints, it is not covered in this review.
2.3.1. Apoptosis
For the past two decade or so, mitochondria have been extensively studied for its essential role in defining the balance between cell life and death. It was first demonstrated in 1994, that cytochrome c, once it is released from the intermembrane space of the mitochondria, can initiate an enzyme cascade of cellular self-destruction. Other death amplification factors, such as the inhibitors of apoptosis (the IAP family which prevent accidental caspase activation presumably), smac or diablo (which inhibits IAPs to permit the apoptotic cascade to proceed), procaspase-9, and AIF (the flavoprotein apoptosis inducing factor) are also released to participate in the death execution pathway [97].
2.3.2. Redox Status in Cell Fate Decision
Although overwhelming amount of ROS are definitely detrimental to the cells, emerging evidence has demonstrated that when they are present in nonlethal concentrations, they can function as proliferative and/or survival signals [8]. A mild increase in the O2 ∙− has been shown to confer survival advantage to the tumor cells under apoptotic triggers [8, 10, 100–103]. Furthermore, it has been demonstrated by our group that cell fate is tightly regulated as a function of the ratio of O2 ∙− and H2O2. A tilt in the balance of the two reactive oxygen species towards O2 ∙− leads to survival signaling, while the reverse sensitizes cells to apoptotic triggers [8–10]. Corroborating the survival advantage of a mild prooxidant status, the antiapoptotic activity of Bcl-2 can be contributed to its noncanonical ability to induce intramitochondrial O2 ∙− production. Indeed, Bcl-2 BH3 peptides, which reversed the prooxidant state of Bcl-2-overexpressing tumor cells, sensitized those cells to drug-induced apoptosis. Similarly, silencing or pharmacological inhibition of Rac1 also compromised Bcl-2-induced increase in intramitochondrial O2 ∙− levels leading to sensitization of those tumor cells to apoptotic triggers [24].
3. Mitochondrial Dysfunction in Pathology and Therapeutic Targeting
Mitochondria, as one of the major ROS producers within the cell, have been rendered susceptible to oxidative damage when the antioxidant defense machinery fails to meet their ROS scavenging tasks; therefore, they are implicated in the pathology of various diseases including neurodegenerative diseases, diabetes, cardiovascular diseases, and cancer [104–114].
3.1. Neurodegenerative Diseases
ROS-mediated mitochondrial dysfunctions and apoptosis have been demonstrated as causal factors in the pathology of several neurodegenerative diseases including Parkinson's disease (PD), Alzheimer's disease (AD), and Amyotrophic lateral sclerosis (ALS). Oxidative damage, as indicated by malondialdehyde (MDA) and 4-hydroxynonenal (4-HNE), have been identified in patients diagnosed with PD (substantia nigra), AD (hippocampus and cortex) as well as ALS (spinal fluid) [115–118]. The levels of Iron are also found to be elevated in the substantia nigra of patients with PD, which could serve as a catalyst for the Fenton's reaction in producing hydroxyl radicals [119–121], whereas the activities of antioxidant defense enzymes such as glutathione peroxidase (GPx) and reductase (GR), superoxide dismutase (SOD) and catalase (CAT) are reduced in the affected brain regions of patients with AD [117, 122, 123]. Similarly, patients with PD demonstrate diminished levels of GSH in the dopaminergic neurons of substantia nigra [118, 119, 124].
Transgenic animals that develop neurodegenerative diseases have been utilized as models to clarify the roles of oxidative stress in the pathogenesis of these diseases. Transgenic mice that harbor an ALS-linked mutant CuZnSOD gene show progressive accumulation of 8-OHdG, one of the best markers for oxidative DNA damage, in ventral horn neurons. The immunoreactivity for this marker indicates the existence of oxidative damage to mitochondrial DNA in spinal motoneurons starting from very early stage of the disease, and probably contributing to the subsequent motoneuron death [125]. In another model of GPx deficient mice, administration of N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) led to enhanced toxicity to dopaminergic neurons indicating the involvement of ROS in the early pathogenesis of PD [126]. Triple-transgenic mice that mimic AD progression in humans also demonstrate reduced levels of GSH and vitamin E as well as increased extent of lipid peroxidation during the stage of Amyloid β (Aβ) oligomerization before the onset of Aβ plaques and neurofibrillary tangles [127].
Since oxidative stress resulting from depletion of the cellular antioxidant glutathione occured during the early stages of neurodegeneration, a lot of effort has been dedicated to study the effects of antioxidants in combating oxidative stress and preserving the integrity of mitochondria for the treatment of neurodegenerative diseases. Several examples of antioxidants that have been demonstrated to attenuate disease progression will be discussed below.
Curcumin —
Curcumin is a polyphenol that belongs to the ginger family (Zingiberaceae). Curcumin treatment of dopaminergic neuronal cells and mice restores GSH pools, which protects against oxidative stress and preserves mitochondrial complex I activity, thereby suggesting its therapeutic benefits in the treatment of PD [128]. Protective effects of curcumin have also been shown against MPP (+)-induced cytotoxicity and apoptosis by upregulation of Bcl-2 expression and restoring mitochondrial membrane potential [129]. In addition, tetrahydrocurcumin (THC), a metabolite of curcumin, has been shown to protect against Aβ-induced ROS burst, preserve mitochondrial membrane potential, and prevent caspase activation in rat primary hippocampal cultures [130].
Apart from oxidative stress, nitrosative stress, largely mediated by reactive nitrogen species (RNS) such as OONO−, is also crucial in PD development by inducing mitochondrial dysfunction through inhibition of brain mitochondrial complex I activity, decrease of mitochondrial membrane potential, and compromise of mitochondrial integrity. The glutamoyl diester bioconjugate of curcumin has been shown to restore Complex I activity and protect against protein nitration [131].
Epigallocatechin-3-Gallate (EGCG) —
EGCG is the most abundant polyphenol found in green tea. Apart from its iron-chelating property, the antioxidant capacity of EGCG has been demonstrated at the level of mitochondria where it not only enhances the activities of both TCA cycle enzymes and ETC complexes but also upregulates the antioxidant system in aged brain [132]. EGCG has also been shown to increase the activity of SOD and catalase in mice striatum [133]. The molecular mechanism underlying the upregulation of the antioxidant defense machinery is due to the ability of EGCG to induce the activation of Nrf2, a master transcription factor for antioxidant and phase II detoxifying enzymes [134].
Although these molecules demonstrate protective effects in animal models of the neurodegenerative diseases, their beneficial effects in humans have not been clearly demonstrated in clinical trials, conceivably due to the difficulties in penetrating the blood-brain barrier; therefore a lot of effort is currently being spent in developing better delivery systems for specific targeting, especially to mitochondria, where their pharmacological activity is mostly required to increase the therapeutic efficacy [135–138].
3.2. Diabetes, Diabetic Complications, and Cardiovascular Diseases
Despite the fact that diabetes mellitus (DM) is a heterogeneous, multifactorial and chronic disease, it can be stated without doubt that DM is linked to acute and continuous overproduction of ROS and characterized by mitochondrial impairment. DM is also marked with chronic inflammation that further weakens intracellular antioxidant defense. Reduced levels of antioxidants such as GSH, vitamins C and E are observed in diabetic patients [139, 140]. Dysfunction of mitochondrial complex I and subsequent increase in ROS production together with decreases in antioxidant levels and membrane potential have been reported in diabetic patients [141]. Risk factors such as aging, obesity, and unhealthy diet contribute to an oxidative environment which impairs insulin signaling and mitochondrial function leading to diabetes development, and the resulting hyperglycemia in turn contributes to the maintenance and progression of the overall oxidative stress through mechanisms such as glycation of antioxidant enzymes [111] and overproduction of O2 − by the mitochondrial ETC, which in turn activate a variety of proinflammatory signals [111, 142–144].
Malfunctioning of mitochondrial oxidative phosphorylation has been considered as one of the main culprits in the development of diabetic complications as well, such as renal dysfunction [145]. In addition, it has been reported that the activities and expression of antioxidant enzymes are decreased in diabetic microvascular disease [146, 147] and a specific polymorphic MnSOD gene is correlated with diabetic nephropathy development [148]. Protective effects of catalase overexpression have also been demonstrated in the experimental models of type 2 diabetic nephropathy, thereby implicating H2O2 [149]. Majority of the type 2 diabetic patients with insulin resistance also exhibit a significantly higher risk of developing cardiovascular disease (CVD) [150]. Human atherosclerotic samples show higher extent of mitochondrial DNA damage, which correlates with greater ROS production. In apoE-null mice, mitochondrial damage has been shown to precede the development of atherosclerosis. In the same model, heterozygous deficiency of MnSOD also increases vascular mitochondrial dysfunction [151].
Since oxidative stress has been considered as one of the main contributing factors for the onset and progression of diabetes, diabetic complications, and CVD, the need to eradicate ROS especially from mitochondria is of great therapeutic importance [152]. Although classical antioxidants such as vitamins C and E do not show significant improvement in disease conditions [153], recent reports show that a subgroup of type 2 diabetic patients with haptoglobin (Hp) 2-2 genotype can benefit from vitamin E supplementation suggesting that tailored treatment regimens for different subgroups of patients could be more favorable [154, 155]. MitoQ is an antioxidant that is selectively targeted to and accumulates in mitochondria due to its covalent attachment to the lipophilic triphenylphosphonium cation [156]. MitoQ administration to Ins2(+/) (AkitaJ) mice improves the tubular and glomerular function and reduces urinary albumin levels and interstitial fibrosis implicating their therapeutic benefits in treating diabetic nephropathy [157].
3.3. Cancer
3.3.1. Prooxidant Theory of Carcinogenesis
We have discussed in the previous Section 2.3.2 how redox status can affect cell fate decision and how some mitochondrial proteins can act as redox mediators. Over the past decade, our group has been working on the underlying mechanisms and translational relevance of redox signaling in the context of carcinogenesis. We have established that cell fate is tightly regulated as a function of the ratio of O2 ∙− and H2O2. A tilt in the balance of the two reactive oxygen species towards O2 ∙− leads to survival signaling while the reverse sensitizes cells to apoptotic triggers [8–10].
3.3.2. Mitochondria as Therapeutic Targets in Cancer
Since mitochondria are key regulators for energy metabolism, ROS production and cell fate [158], targeting mitochondria to elicit cell death would therefore be a good strategy in cancer therapeutics especially in those cancer cells where upstream apoptotic signaling is malfunctional [159]. In addition, the bioenergetic differences between nontransformed and cancer cells confer the selectivity and specificity in targeting small molecule agents to the mitochondria of desired cancer cells. For example, lipophilic cations preferentially accumulate in the mitochondrial matrix of cancer cells due to their increased mitochondrial membrane potential resulting from increased glycolytic rates as compared to their normal counterparts [160, 161]. In addition, the addiction of cancer cells to glycolysis for ATP supply, described as the Warburg effect, also renders them more susceptible to apoptotic induction when their cellular bioenergetic pathways are intervened [161].
Direct Targeting of Mitochondrial ETC —
Since mitochondrial ETC is essential in energy production and ROS generation, there are a plethora of agents that target ETC directly for cancer therapeutics [106]. Rotenone is a naturally derived hydrophobic pesticide, which binds and inactivates mitochondrial Complex I irreversibly leading to a blockage of oxidative phosphorylation and an increase in ROS generation [162, 163] and finally apoptotic induction [164–167]. Therapeutic benefits of Rotenone have been demonstrated in human breast cancer cells [168], neuroblastomas [169], promyelocytic leukemias [170], and human B-cell lymphomas [167]. Tamoxifen and estradiol have also been reported to act on the flavin mononucleotide site of complex I leading to mitochondrial failure independent of estrogen receptors [171], and particularly in MCF-7 breast cancer cells tamoxifen is able to induce an increase in ROS levels, a decrease in mitochondrial membrane potential and release of cytochrome c [172]. 3-nitropropionic acid is a toxin found in fungi and plants that can bind covalently to complex II [173–175] and its toxicity to tumor cells is linked to cellular energy depletion and oxidative stress due to the generation of O2 −, H2O2, and OONO− [176, 177]. In addition, an analog of vitamin E, α-tocopheryl succinate (α-TOS) is able to interfere with the ubiquinone binding site on Complex II [178] leading to cell cycle arrest and apoptosis in a host of established cancer cell lines of different origin as well as in in vivo experimental animal models [179–185]. Although α-TOS has been shown to be largely nontoxic to normal tissues [183], its efficiency has not yet been tested in human cancer patients due to the difficulties in administration [186]. Antimycin A is a secondary metabolite produced by Streptomyces kitazawensis [187] which binds to the Qi site of Complex III [188, 189] leading to collapse of the proton gradient [188], ROS production, and apoptosis [190]. Fenretinide is a synthetic analog of retinoic acid which is able to downregulate Complex IV subunit III mRNA levels leading to a decrease in Complex IV activity [191]. Fenretinide induces apoptosis through elevated ROS production, cytochrome c release, and induction of mitochondrial permeability transition [192, 193], which could be prevented by the addition of antioxidants [194, 195]. It is likely that fenretinide inhibits at least one of the complexes along the ETC, although the exact prooxidant mechanisms are yet to be elucidated [194, 196]. In vivo therapeutic benefits of fenretinide have been demonstrated in both carcinogen-induced or xenograft animal models [197].
Direct targeting of mitochondrial ETC increases ROS production from the mitochondria of cancer cells which results in increased susceptibility of those glycolytic addicted cells to apoptotic induction. However, the critical points to be taken into consideration when using mitochondrial respiration “poisons” are their in vivo toxicity and therapeutic indexes. Almost half of the studies as discussed above failed to actually demonstrate nontoxicity of the agents to nontransformed cells in vivo except for α-TOS and fenretinide.
Direct Targeting of Bcl-2 Family Proteins —
Apart from targeting the mitochondrial ETC, there is another group of proteins that are of particular interest due to their regulatory roles in apoptosis, which is the Bcl-2 family. We have discussed in Sections 2.2.2 and 2.3.2 that the ratio between the pro- and antiapoptotic members of Bcl-2 family is critical in cell fate decision. In addition, Bcl-2, an antiapoptotic member of the family and a resident protein of mitochondria, is able to modulate redox status which could be utilized in cancer therapeutics as well. Promising therapeutic strategies that aim at overcoming the problem of Bcl-2 overexpression (which happens in a number of cancers) including Bcl-2 antisense and BH3 mimetics have been recently reviewed by our group [66]. Furthermore, treatment of membrane active segments of the proapoptotic member Bax can also induce apoptosis in tumor cells [198].
Indirect Targeting of Mitochondrial Apoptotic Pathway —
There are another group of drugs that do not target mitochondria directly but rather modulate mitochondrial proteins and/or induce ROS production leading to induction of intrinsic mitochondrial apoptotic pathway in cancer cells. These include clinically used chemotherapies such as irinotecan, topotecan [199], etoposide [200], vinblastine [201], and arsenic trioxide [202], which induce mitochondrial apoptotic pathway as well as those currently undergoing clinical evaluation such as betulinic acid [203], curcumin [204], camptothecin derivatives [199], and triapine [205], where betulinic acid and triapine-induced apoptosis has been attributed to ROS production.
4. Concluding Remarks
Mitochondria are essential regulators of cellular energy metabolism, redox homeostasis, and cell fate decision, and their dysfunction inevitably leads to various pathological states including neurodegenerative diseases, diabetes, cardiovascular diseases, and cancer as briefly discussed in this review. Oxidative stress is the underlying causal factor in majority if not all of the diseases listed above; therefore, therapeutic strategies that aim at manipulating the redox metabolism represent promising options which have been and will still be at the center stage of targeted drug development.
References
- 1.Duchen MR. Mitochondria in health and disease: perspectives on a new mitochondrial biology. Molecular Aspects of Medicine. 2004;25(4):365–451. doi: 10.1016/j.mam.2004.03.001. [DOI] [PubMed] [Google Scholar]
- 2.Hajnóczky G, Csordás G, Das S, et al. Mitochondrial calcium signalling and cell death: approaches for assessing the role of mitochondrial Ca2+ uptake in apoptosis. Cell Calcium. 2006;40(5-6):553–560. doi: 10.1016/j.ceca.2006.08.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Oh-Hama T. Evolutionary consideration on 5-aminolevulinate synthase in nature. Origins of Life and Evolution of the Biosphere. 1997;27(4):405–412. doi: 10.1023/a:1006583601341. [DOI] [PubMed] [Google Scholar]
- 4.Rossier MF. T channels and steroid biosynthesis: in search of a link with mitochondria. Cell Calcium. 2006;40(2):155–164. doi: 10.1016/j.ceca.2006.04.020. [DOI] [PubMed] [Google Scholar]
- 5.McBride HM, Neuspiel M, Wasiak S. Mitochondria: more than just a powerhouse. Current Biology. 2006;16(14):R551–R560. doi: 10.1016/j.cub.2006.06.054. [DOI] [PubMed] [Google Scholar]
- 6.Green DR. Apoptotic pathways: the roads to ruin. Cell. 1998;94(6):695–698. doi: 10.1016/s0092-8674(00)81728-6. [DOI] [PubMed] [Google Scholar]
- 7.Hamanaka RB, Chandel NS. Mitochondrial reactive oxygen species regulate cellular signaling and dictate biological outcomes. Trends in Biochemical Sciences. 2010;35(9):505–513. doi: 10.1016/j.tibs.2010.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Pervaiz S, Clement MV. Superoxide anion: oncogenic reactive oxygen species? International Journal of Biochemistry and Cell Biology. 2007;39(7-8):1297–1304. doi: 10.1016/j.biocel.2007.04.007. [DOI] [PubMed] [Google Scholar]
- 9.Clément MV, Pervaiz S. Intracellular superoxide and hydrogen peroxide concentrations: a critical balance that determines survival or death. Redox Report. 2001;6(4):211–214. doi: 10.1179/135100001101536346. [DOI] [PubMed] [Google Scholar]
- 10.Clément MV, Ponton A, Pervaiz S. Apoptosis induced by hydrogen peroxide is mediated by decreased superoxide anion concentration and reduction of intracellular milieu. FEBS Letters. 1998;440(1-2):13–18. doi: 10.1016/s0014-5793(98)01410-0. [DOI] [PubMed] [Google Scholar]
- 11.Leblanc C, Richard O, Kloareg B, Viehmann S, Zetsche K, Boyen C. Origin and evolution of mitochondria: what have we learnt from red algae? Current Genetics. 1997;31(3):193–207. doi: 10.1007/s002940050196. [DOI] [PubMed] [Google Scholar]
- 12.Stock D, Leslie AGW, Walker JE. Molecular architecture of the rotary motor in ATP synthase. Science. 1999;286(5445):1700–1705. doi: 10.1126/science.286.5445.1700. [DOI] [PubMed] [Google Scholar]
- 13.Cadenas E, Boveris A, Ragan CI, Stoppani AOM. Production of superoxide radicals and hydrogen peroxide by NADH ubiquinone reductase and ubiquinol cytochrome c reductase from beef heart mitochondria. Archives of Biochemistry and Biophysics. 1977;180(2):248–257. doi: 10.1016/0003-9861(77)90035-2. [DOI] [PubMed] [Google Scholar]
- 14.Han D, Williams E, Cadenas E. Mitochondrial respiratory chain-dependent generation of superoxide anion and its release into the intermembrane space. Biochemical Journal. 2001;353, part 2:411–416. doi: 10.1042/0264-6021:3530411. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.St-Pierre J, Buckingham JA, Roebuck SJ, Brand MD. Topology of superoxide production from different sites in the mitochondrial electron transport chain. Journal of Biological Chemistry. 2002;277(47):44784–44790. doi: 10.1074/jbc.M207217200. [DOI] [PubMed] [Google Scholar]
- 16.Tahara EB, Navarete FDT, Kowaltowski AJ. Tissue-, substrate-, and site-specific characteristics of mitochondrial reactive oxygen species generation. Free Radical Biology and Medicine. 2009;46(9):1283–1297. doi: 10.1016/j.freeradbiomed.2009.02.008. [DOI] [PubMed] [Google Scholar]
- 17.Piskernik C, Haindl S, Behling T, et al. Antimycin A and lipopolysaccharide cause the leakage of superoxide radicals from rat liver mitochondria. Biochimica et Biophysica Acta. 2008;1782(4):280–285. doi: 10.1016/j.bbadis.2008.01.007. [DOI] [PubMed] [Google Scholar]
- 18.Brand MD. The sites and topology of mitochondrial superoxide production. Experimental Gerontology. 2010;45(7-8):466–472. doi: 10.1016/j.exger.2010.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Ježek P, Hlavatá L. Mitochondria in homeostasis of reactive oxygen species in cell, tissues, and organism. International Journal of Biochemistry and Cell Biology. 2005;37(12):2478–2503. doi: 10.1016/j.biocel.2005.05.013. [DOI] [PubMed] [Google Scholar]
- 20.Li JM, Shah AM. ROS generation by nonphagocytic NADPH oxidase: potential relevance in diabetic nephropathy. Journal of the American Society of Nephrology. 2003;14(8, supplement 3):S221–S226. doi: 10.1097/01.asn.0000077406.67663.e7. [DOI] [PubMed] [Google Scholar]
- 21.Block K, Gorin Y, Abboud HE. Subcellular localization of Nox4 and regulation in diabetes. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(34):14385–14390. doi: 10.1073/pnas.0906805106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Kuroda J, Ago T, Matsushima S, Zhai P, Schneider MD, Sadoshima J. NADPH oxidase 4 (Nox4) is a major source of oxidative stress in the failing heart. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(35):15565–15570. doi: 10.1073/pnas.1002178107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Ago T, Kuroda J, Pain J, Fu C, Li H, Sadoshima J. Upregulation of Nox4 by hypertrophic stimuli promotes apoptosis and mitochondrial dysfunction in cardiac myocytes. Circulation Research. 2010;106(7):1253–1264. doi: 10.1161/CIRCRESAHA.109.213116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Velaithan R, Kang J, Hirpara JL, et al. The small GTPase Rac1 is a novel binding partner of Bcl-2 and stabilizes its antiapoptotic activity. Blood. 2011;117(23):6214–6226. doi: 10.1182/blood-2010-08-301283. [DOI] [PubMed] [Google Scholar]
- 25.Doughan AK, Harrison DG, Dikalov SI. Molecular mechanisms of angiotensin II-mediated mitochondrial dysfunction: linking mitochondrial oxidative damage and vascular endothelial dysfunction. Circulation Research. 2008;102(4):488–496. doi: 10.1161/CIRCRESAHA.107.162800. [DOI] [PubMed] [Google Scholar]
- 26.Starkov AA, Fiskum G, Chinopoulos C, et al. Mitochondrial α-ketoglutarate dehydrogenase complex generates reactive oxygen species. Journal of Neuroscience. 2004;24(36):7779–7788. doi: 10.1523/JNEUROSCI.1899-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Fang H, Chen M, Ding Y, et al. Imaging superoxide flash and metabolism-coupled mitochondrial permeability transition in living animals. Cell Research. 2011;21(9):1295–1304. doi: 10.1038/cr.2011.81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Wang W, Fang H, Groom L, et al. Superoxide Flashes in Single Mitochondria. Cell. 2008;134(2):279–290. doi: 10.1016/j.cell.2008.06.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Mattson MP. A reaction to mitochondria in action. Cell Research. 2011;21(9):1279–1282. [Google Scholar]
- 30.Ghafourifar P, Richter C. Nitric oxide synthase activity in mitochondria. FEBS Letters. 1997;418(3):291–296. doi: 10.1016/s0014-5793(97)01397-5. [DOI] [PubMed] [Google Scholar]
- 31.Dedkova EN, Ji X, Lipsius SL, Blatter LA. Mitochondrial calcium uptake stimulates nitric oxide production in mitochondria of bovine vascular endothelial cells. American Journal of Physiology. 2004;286(2):C406–C415. doi: 10.1152/ajpcell.00155.2003. [DOI] [PubMed] [Google Scholar]
- 32.Rakhit RD, Mojet MH, Marber MS, Duchen MR. Mitochondria as targets for nitric oxide-induced protection during simulated ischemia and reoxygenation in isolated neonatal cardiomyocytes. Circulation. 2001;103(21):2617–2623. doi: 10.1161/01.cir.103.21.2617. [DOI] [PubMed] [Google Scholar]
- 33.Poderoso JJ, Carreras MC, Lisdero C, Riobó N, Schöpfer F, Boveris A. Nitric oxide inhibits electron transfer and increases superoxide radical production in rat heart mitochondria and submitochondrial particles. Archives of Biochemistry and Biophysics. 1996;328(1):85–92. doi: 10.1006/abbi.1996.0146. [DOI] [PubMed] [Google Scholar]
- 34.Cassina A, Radi R. Differential inhibitory action of nitric oxide and peroxynitrite on mitochondrial electron transport. Archives of Biochemistry and Biophysics. 1996;328(2):309–316. doi: 10.1006/abbi.1996.0178. [DOI] [PubMed] [Google Scholar]
- 35.Cleeter MW, Cooper JM, Darley-Usmar VM, Moncada S, Schapira AH. Reversible inhibition of cytochrome c oxidase, the terminal enzyme of the mitochondrial respiratory chain, by nitric oxide Implications for neurodegenerative diseases. FEBS Letters. 1994;345(1):50–54. doi: 10.1016/0014-5793(94)00424-2. [DOI] [PubMed] [Google Scholar]
- 36.Brown GC. Regulation of mitochondrial respiration by nitric oxide inhibition of cytochrome c oxidase. Biochimica et Biophysica Acta. 2001;1504(1):46–57. doi: 10.1016/s0005-2728(00)00238-3. [DOI] [PubMed] [Google Scholar]
- 37.Borutaité V, Brown GC. Rapid reduction of nitric oxide by mitochondria, and reversible inhibition of mitochondrial respiration by nitric oxide. Biochemical Journal. 1996;315(1):295–299. doi: 10.1042/bj3150295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Takehara Y, Kanno T, Yoshioka T, Inoue M, Utsumi K. Oxygen-dependent regulation of mitochondrial energy metabolism by nitric oxide. Archives of Biochemistry and Biophysics. 1995;323(1):27–32. doi: 10.1006/abbi.1995.0005. [DOI] [PubMed] [Google Scholar]
- 39.Torres J, Darley-Usmar V, Wilson MT. Inhibition of cytochrome c oxidase in turnover by nitric oxide: mechanism and implications for control of respiration. Biochemical Journal. 1995;312(1):169–173. doi: 10.1042/bj3120169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Brown GC. Nanomolar concentrations of nitric oxide reversibly inhibit synaptosomal respiration by competing with oxygen at cytochrome oxidase. FEBS Letters. 1994;356(2-3):295–298. doi: 10.1016/0014-5793(94)01290-3. [DOI] [PubMed] [Google Scholar]
- 41.Nisoli E, Clementi E, Paolucci C, et al. Mitochondrial biogenesis in mammals: the role of endogenous nitric oxide. Science. 2003;299(5608):896–899. doi: 10.1126/science.1079368. [DOI] [PubMed] [Google Scholar]
- 42.Radi R, Cassina A, Hodara R, Quijano C, Castro L. Peroxynitrite reactions and formation in mitochondria. Free Radical Biology and Medicine. 2002;33(11):1451–1464. doi: 10.1016/s0891-5849(02)01111-5. [DOI] [PubMed] [Google Scholar]
- 43.Burwell LS, Nadtochiy SM, Brookes PS. Cardioprotection by metabolic shut-down and gradual wake-up. Journal of Molecular and Cellular Cardiology. 2009;46(6):804–810. doi: 10.1016/j.yjmcc.2009.02.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Radi R, Rodriguez M, Castro L, Telleri R. Inhibition of mitochondrial electron transport by peroxynitrite. Archives of Biochemistry and Biophysics. 1994;308(1):89–95. doi: 10.1006/abbi.1994.1013. [DOI] [PubMed] [Google Scholar]
- 45.Sharpe MA, Cooper CE. Reactions of nitric oxide with mitochondrial cytochrome c: a novel mechanism for the formation of nitroxyl anion and peroxynitrite. Biochemical Journal. 1998;332(1):9–19. doi: 10.1042/bj3320009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Poderoso JJ, Carreras MC, Schöpfer F, et al. The reaction of nitric oxide with ubiquinol: kinetic properties and biological significance. Free Radical Biology and Medicine. 1999;26(7-8):925–935. doi: 10.1016/s0891-5849(98)00277-9. [DOI] [PubMed] [Google Scholar]
- 47.Clementi E, Brown GC, Feelisch M, Moncada S. Persistent inhibition of cell respiration by nitric oxide: crucial role of S-nitrosylation of mitochondrial complex I and protective action of glutathione. Proceedings of the National Academy of Sciences of the United States of America. 1998;95(13):7631–7636. doi: 10.1073/pnas.95.13.7631. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Riobó NA, Clementi E, Melani M, et al. Nitric oxide inhibits mitochondrial NADH:ubiquinone reductase activity through peroxynitrite formation. Biochemical Journal. 2001;359(1):139–145. doi: 10.1042/0264-6021:3590139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Borutaite V, Budriunaite A, Brown GC. Reversal of nitric oxide-, peroxynitrite- and S-nitrosothiol-induced inhibition of mitochondrial respiration or complex I activity by light and thiols. Biochimica et Biophysica Acta. 2000;1459(2-3):405–412. doi: 10.1016/s0005-2728(00)00178-x. [DOI] [PubMed] [Google Scholar]
- 50.Szabó C, Salzman AL. Endogenous peroxynitrite is involved in the inhibition of mitochondrial respiration in immune-stimulated J774.2 macrophages. Biochemical and Biophysical Research Communications. 1995;209(2):739–743. doi: 10.1006/bbrc.1995.1561. [DOI] [PubMed] [Google Scholar]
- 51.Bolanos JP, Heales SJR, Land JM, Clark JB. Effect of peroxynitrite on the mitochondrial respiratory chain: differential susceptibility of neurones and astrocytes in primary culture. Journal of Neurochemistry. 1995;64(5):1965–1972. doi: 10.1046/j.1471-4159.1995.64051965.x. [DOI] [PubMed] [Google Scholar]
- 52.Lizasoain I, Moro MA, Knowles RG, Darley-Usmar V, Moncada S. Nitric oxide and peroxynitrite exert distinct effects on mitochondrial respiration which are differentially blocked by glutathione or glucose. Biochemical Journal. 1996;314(3):877–880. doi: 10.1042/bj3140877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Xie YW, Wolin MS. Role of nitric oxide and its interaction with superoxide in the suppression of cardiac muscle mitochondrial respiration: involvement in response to hypoxia/reoxygenation. Circulation. 1996;94(10):2580–2586. doi: 10.1161/01.cir.94.10.2580. [DOI] [PubMed] [Google Scholar]
- 54.Quijano C, Hernandez-Saavedra D, Castro L, McCord JM, Freeman BA, Radi R. Reaction of peroxynitrite with Mn-Superoxide dismutase. Role of the metal center in decomposition kinetics and nitration. Journal of Biological Chemistry. 2001;276(15):11631–11638. doi: 10.1074/jbc.M009429200. [DOI] [PubMed] [Google Scholar]
- 55.Vieira HLA, Belzacq AS, Haouzi D, et al. The adenine nucleotide translocator: a target of nitric oxide, peroxynitrite, and 4-hydroxynonenal. Oncogene. 2001;20(32):4305–4316. doi: 10.1038/sj.onc.1204575. [DOI] [PubMed] [Google Scholar]
- 56.Konorev EA, Hogg N, Kalyanaraman B. Rapid and irreversible inhibition of creatine kinase by peroxynitrite. FEBS Letters. 1998;427(2):171–174. doi: 10.1016/s0014-5793(98)00413-x. [DOI] [PubMed] [Google Scholar]
- 57.Forsmark-Andrée P, Persson B, Radi R, Dallner G, Ernster L. Oxidative modification of nicotinamide nucleotide transhydrogenase in submitochondrial particles: effect of endogenous ubiquinol. Archives of Biochemistry and Biophysics. 1996;336(1):113–120. doi: 10.1006/abbi.1996.0538. [DOI] [PubMed] [Google Scholar]
- 58.Castro L, Rodriguez M, Radi R. Aconitase is readily inactivated by peroxynitrite, but not by its precursor, nitric oxide. Journal of Biological Chemistry. 1994;269(47):29409–29415. [PubMed] [Google Scholar]
- 59.Schweizer M, Richter C. Peroxynitrite stimulates the pyridine nucleotide-linked Ca2+ release from intact rat liver mitochondria. Biochemistry. 1996;35(14):4524–4528. doi: 10.1021/bi952708+. [DOI] [PubMed] [Google Scholar]
- 60.Kuzkaya N, Weissmann N, Harrison DG, Dikalov S. Interactions of peroxynitrite, tetrahydrobiopterin, ascorbic acid, and thiols: implications for uncoupling endothelial nitric-oxide synthase. Journal of Biological Chemistry. 2003;278(25):22546–22554. doi: 10.1074/jbc.M302227200. [DOI] [PubMed] [Google Scholar]
- 61.Piantadosi CA. Carbon monoxide, reactive oxygen signaling, and oxidative stress. Free Radical Biology and Medicine. 2008;45(5):562–569. doi: 10.1016/j.freeradbiomed.2008.05.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Patenaude A, Ven Murthy MR, Mirault ME. Mitochondrial thioredoxin system: effects of TrxR2 overexpression on redox balance, cell growth, and apoptosis. Journal of Biological Chemistry. 2004;279(26):27302–27314. doi: 10.1074/jbc.M402496200. [DOI] [PubMed] [Google Scholar]
- 63.Naik E, Dixit VM. Mitochondrial reactive oxygen species drive proinflammatory cytokine production. Journal of Experimental Medicine. 2011;208(3):417–420. doi: 10.1084/jem.20110367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Hwang AB, Lee SJ. Regulation of life span by mitochondrial respiration: the HIF-1 and ROS connection. Aging. 2011;3(3):304–310. doi: 10.18632/aging.100292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Krajewski S, Tanaka S, Takayama S, Schibler MJ, Fenton W, Reed JC. Investigation of the subcellular distribution of the bcl-2 oncoprotein: residence in the nuclear envelope, endoplasmic reticulum, and outer mitochondrial membranes. Cancer Research. 1993;53(18):4701–4714. [PubMed] [Google Scholar]
- 66.Low ICC, Kang J, Pervaiz S. Bcl-2: a prime regulator of mitochondrial redox metabolism in cancer cells. Antioxidants and Redox Signaling. 2011;15(12):2975–2987. doi: 10.1089/ars.2010.3851. [DOI] [PubMed] [Google Scholar]
- 67.Hockenbery DM, Oltvai ZN, Yin XM, Milliman CL, Korsmeyer SJ. Bcl-2 functions in an antioxidant pathway to prevent apoptosis. Cell. 1993;75(2):241–251. doi: 10.1016/0092-8674(93)80066-n. [DOI] [PubMed] [Google Scholar]
- 68.Chen SR, Dunigan DD, Dickman MB. Bcl-2 family members inhibit oxidative stress-induced programmed cell death in Saccharomyces cerevisiae. Free Radical Biology and Medicine. 2003;34(10):1315–1325. doi: 10.1016/s0891-5849(03)00146-1. [DOI] [PubMed] [Google Scholar]
- 69.Kane DJ, Sarafian TA, Anton R, et al. Bcl-2 inhibition of neural death: decreased generation of reactive oxygen species. Science. 1993;262(5137):1274–1277. doi: 10.1126/science.8235659. [DOI] [PubMed] [Google Scholar]
- 70.Mirkovic N, Voehringer DW, Story MD, McConkey DJ, McDonnell TJ, Meyn RE. Resistance to radiation-induced apoptosis in Bcl-2-expressing cells is reversed by depleting cellular thiols. Oncogene. 1997;15(12):1461–1470. doi: 10.1038/sj.onc.1201310. [DOI] [PubMed] [Google Scholar]
- 71.Myers KM, Fiskum G, Liu Y, Simmens SJ, Bredesen DE, Murphy AN. Bcl-2 protects neural cells from cyanide/aglycemia-induced lipid oxidation, mitochondrial injury, and loss of viability. Journal of Neurochemistry. 1995;65(6):2432–2440. doi: 10.1046/j.1471-4159.1995.65062432.x. [DOI] [PubMed] [Google Scholar]
- 72.Zamzami N, Marzo I, Susin SA, et al. The thiol crosslinking agent diamide overcomes the apoptosis-inhibitory effect of Bcl-2 by enforcing mitochondrial permeability transition. Oncogene. 1998;16(8):1055–1063. doi: 10.1038/sj.onc.1201864. [DOI] [PubMed] [Google Scholar]
- 73.Zhong LT, Sarafian T, Kane DJ, et al. bcl-2 inhibits death of central neural cells induced by multiple agents. Proceedings of the National Academy of Sciences of the United States of America. 1993;90(10):4533–4537. doi: 10.1073/pnas.90.10.4533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Lee M, Hyun DH, Marshall KA, et al. Effect of overexpression of Bcl-2 on cellular oxidative damage, nitric oxide production, antioxidant defenses, and the proteasome. Free Radical Biology and Medicine. 2001;31(12):1550–1559. doi: 10.1016/s0891-5849(01)00633-5. [DOI] [PubMed] [Google Scholar]
- 75.Chen ZX, Pervaiz S. BCL-2: pro-or anti-oxidant? Frontiers in Bioscience. 2009;1:263–268. doi: 10.2741/E25. [DOI] [PubMed] [Google Scholar]
- 76.Ellerby LM, Ellerby HM, Park SM, et al. Shift of the cellular oxidation-reduction potential in neural cells expressing Bcl-2. Journal of Neurochemistry. 1996;67(3):1259–1267. doi: 10.1046/j.1471-4159.1996.67031259.x. [DOI] [PubMed] [Google Scholar]
- 77.Kowaltowski AJ, Fiskum G. Redox mechanisms of cytoprotection by Bcl-2. Antioxidants and Redox Signaling. 2005;7(3-4):508–514. doi: 10.1089/ars.2005.7.508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Kowaltowski AJ, Vercesi AE, Fiskum G. Bcl-2 prevents mitochondrial permeability transition and cytochrome c release via maintenance of reduced pyridine nucleotides. Cell Death and Differentiation. 2000;7(10):903–910. doi: 10.1038/sj.cdd.4400722. [DOI] [PubMed] [Google Scholar]
- 79.Papadopoulos MC, Koumenis IL, Xu L, Giffard RG. Potentiation of murine astrocyte antioxidant defence by bcl-2: protection in part reflects elevated glutathione levels. European Journal of Neuroscience. 1998;10(4):1252–1260. doi: 10.1046/j.1460-9568.1998.00134.x. [DOI] [PubMed] [Google Scholar]
- 80.Armstrong JS, Jones DP. Glutathione depletion enforces the mitochondrial permeability transition and causes cell death in Bcl-2 overexpressing HL60 cells. The FASEB Journal. 2002;16(10):1263–1265. doi: 10.1096/fj.02-0097fje. [DOI] [PubMed] [Google Scholar]
- 81.Chen ZX, Pervaiz S. Bcl-2 induces pro-oxidant state by engaging mitochondrial respiration in tumor cells. Cell Death and Differentiation. 2007;14(9):1617–1627. doi: 10.1038/sj.cdd.4402165. [DOI] [PubMed] [Google Scholar]
- 82.Clément MV, Hirpara JL, Pervaiz S. Decrease in intracellular superoxide sensitizes Bcl-2-overexpressing tumor cells to receptor and drug-induced apoptosis independent of the mitochondria. Cell Death and Differentiation. 2003;10(11):1273–1285. doi: 10.1038/sj.cdd.4401302. [DOI] [PubMed] [Google Scholar]
- 83.Esposti MD, Hatzinisiriou I, McLennan H, Ralph S. Bcl-2 and mitochondrial oxygen radicals. New approaches with reactive oxygen species-sensitive probes. Journal of Biological Chemistry. 1999;274(42):29831–29837. doi: 10.1074/jbc.274.42.29831. [DOI] [PubMed] [Google Scholar]
- 84.Steinman HM. The Bcl-2 oncoprotein functions as a pro-oxidant. Journal of Biological Chemistry. 1995;270(8):3487–3490. [PubMed] [Google Scholar]
- 85.Chen ZX, Pervaiz S. Involvement of cytochrome c oxidase subunits Va and Vb in the regulation of cancer cell metabolism by Bcl-2. Cell Death and Differentiation. 2010;17(3):408–420. doi: 10.1038/cdd.2009.132. [DOI] [PubMed] [Google Scholar]
- 86.Kowaltowski AJ, Cosso RG, Campos CB, Fiskum G. Effect of Bcl-2 overexpression on mitochondrial structure and function. Journal of Biological Chemistry. 2002;277(45):42802–42807. doi: 10.1074/jbc.M207765200. [DOI] [PubMed] [Google Scholar]
- 87.Low ICC, Chen ZX, Pervaiz S. Bcl-2 modulates resveratrol-induced ROS production by regulating mitochondrial respiration in tumor cells. Antioxidants and Redox Signaling. 2010;13(6):807–819. doi: 10.1089/ars.2009.3050. [DOI] [PubMed] [Google Scholar]
- 88.Li Campian J, Gao X, Qian M, Eaton JW. Cytochrome c oxidase activity and oxygen tolerance. Journal of Biological Chemistry. 2007;282(17):12430–12438. doi: 10.1074/jbc.M604547200. [DOI] [PubMed] [Google Scholar]
- 89.Lambert AJ, Brand MD. Inhibitors of the quinone-binding site allow rapid superoxide production from mitochondrial NADH:ubiquinone oxidoreductase (complex I) Journal of Biological Chemistry. 2004;279(38):39414–39420. doi: 10.1074/jbc.M406576200. [DOI] [PubMed] [Google Scholar]
- 90.Mann GE, Bonacasa B, Ishii T, Siow RC. Targeting the redox sensitive Nrf2-Keap1 defense pathway in cardiovascular disease: protection afforded by dietary isoflavones. Current Opinion in Pharmacology. 2009;9(2):139–145. doi: 10.1016/j.coph.2008.12.012. [DOI] [PubMed] [Google Scholar]
- 91.Mann GE, Rowlands DJ, Li FYL, de Winter P, Siow RCM. Activation of endothelial nitric oxide synthase by dietary isoflavones: role of NO in Nrf2-mediated antioxidant gene expression. Cardiovascular Research. 2007;75(2):261–274. doi: 10.1016/j.cardiores.2007.04.004. [DOI] [PubMed] [Google Scholar]
- 92.Cheng X, Siow RCM, Mann GE. Impaired redox signaling and antioxidant gene expression in endothelial cells in diabetes: a role for mitochondria and the nuclear factor-E2-related factor 2-Kelch-like ECH-associated protein 1 defense pathway. Antioxidants and Redox Signaling. 2011;14(3):469–487. doi: 10.1089/ars.2010.3283. [DOI] [PubMed] [Google Scholar]
- 93.Niture SK, Jaiswal AK. Nrf2 up-regulates antiapoptotic protein Bcl-2 and prevents cellular apoptosis. The Journal of Biological Chemistry. 2012;287(13):9873–9886. doi: 10.1074/jbc.M111.312694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Kroemer G, Dallaporta B, Resche-Rigon M. The mitochondrial death/life regulator in apoptosis and necrosis. Annual Review of Physiology. 1998;60:619–642. doi: 10.1146/annurev.physiol.60.1.619. [DOI] [PubMed] [Google Scholar]
- 95.Warburg O. On the origin of cancer cells. Science. 1956;123(3191):309–314. doi: 10.1126/science.123.3191.309. [DOI] [PubMed] [Google Scholar]
- 96.Kim JW, Dang CV. Cancer’s molecular sweet tooth and the warburg effect. Cancer Research. 2006;66(18):8927–8930. doi: 10.1158/0008-5472.CAN-06-1501. [DOI] [PubMed] [Google Scholar]
- 97.Zimmermann KC, Bonzon C, Green DR. The machinery of programmed cell death. Pharmacology and Therapeutics. 2001;92(1):57–70. doi: 10.1016/s0163-7258(01)00159-0. [DOI] [PubMed] [Google Scholar]
- 98.Marchi S, Giorgi C, Suski JM, et al. Mitochondria-ros crosstalk in the control of cell death and aging. Journal of Signal Transduction. 2012;2012:17 pages. doi: 10.1155/2012/329635. Article ID 329635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Li Z-Y, Yang Y, Ming M, Liu B. Mitochondrial ROS generation for regulation of autophagic pathways in cancer. Biochemical and Biophysical Research Communications. 2011;414(1):5–8. doi: 10.1016/j.bbrc.2011.09.046. [DOI] [PubMed] [Google Scholar]
- 100.Ahmad KA, Clement MV, Hanif IM, Pervaiz S. Resveratrol Inhibits Drug-Induced Apoptosis in Human Leukemia Cells by Creating an Intracellular Milieu Nonpermissive for Death Execution. Cancer Research. 2004;64(4):1452–1459. doi: 10.1158/0008-5472.can-03-2414. [DOI] [PubMed] [Google Scholar]
- 101.Ahmad KA, Clement MV, Pervaiz S. Pro-oxidant activity of low doses of resveratrol inhibits hydrogen peroxide—induced apoptosis. Annals of the New York Academy of Sciences. 2003;1010:365–373. doi: 10.1196/annals.1299.067. [DOI] [PubMed] [Google Scholar]
- 102.Clément MV, Sivarajah S, Pervaiz S. Production of intracellular superoxide mediates dithiothreitol-dependent inhibition of apoptotic cell death. Antioxidants and Redox Signaling. 2005;7(3-4):456–464. doi: 10.1089/ars.2005.7.456. [DOI] [PubMed] [Google Scholar]
- 103.Pervaiz S, Ramalingam JK, Hirpara JL, Clément MV. Superoxide anion inhibits drug-induced tumor cell death. FEBS Letters. 1999;459(3):343–348. doi: 10.1016/s0014-5793(99)01258-2. [DOI] [PubMed] [Google Scholar]
- 104.Ghosh N, Ghosh R, Mandal SC. Antioxidant protection: a promising therapeutic intervention in neurodegenerative disease. Free Radical Research. 2011;45(8):888–905. doi: 10.3109/10715762.2011.574290. [DOI] [PubMed] [Google Scholar]
- 105.Perrelli MG, Pagliaro P, Penna C. Ischemia/reperfusion injury and cardioprotective mechanisms: role of mitochondria and reactive oxygen species. World Journal of Cardiology. 2011;3(6):186–200. doi: 10.4330/wjc.v3.i6.186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Ferrin G, Linares CI, Muntane J. Mitochondrial drug targets in cell death and cancer. Current Pharmaceutical Design. 2011;17(20):2002–2016. doi: 10.2174/138161211796904803. [DOI] [PubMed] [Google Scholar]
- 107.Victor VM, Rocha M, Bañuls C, Bellod L, Hernandez-Mijares A. Mitochondrial dysfunction and targeted drugs: a focus on diabetes. Current Pharmaceutical Design. 2011;17(20):1986–2001. doi: 10.2174/138161211796904722. [DOI] [PubMed] [Google Scholar]
- 108.Chatterjee A, Dasgupta S, Sidransky D. Mitochondrial subversion in cancer. Cancer Prevention Research. 2011;4(5):638–654. doi: 10.1158/1940-6207.CAPR-10-0326. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Ishii T, Miyazawa M, Hartman PS, Ishii N. Mitochondrial superoxide anion (O2 •−) inducible "mev-1" animal models for aging research. BMB Reports. 2011;44(5):298–305. doi: 10.5483/BMBRep.2011.44.5.298. [DOI] [PubMed] [Google Scholar]
- 110.Al Ghouleh I, Khoo NKH, Knaus UG, et al. Oxidases and peroxidases in cardiovascular and lung disease: new concepts in reactive oxygen species signaling. Free Radical Biology and Medicine. 2011;51(7):1271–1288. doi: 10.1016/j.freeradbiomed.2011.06.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Rains JL, Jain SK. Oxidative stress, insulin signaling, and diabetes. Free Radical Biology and Medicine. 2011;50(5):567–575. doi: 10.1016/j.freeradbiomed.2010.12.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Serviddio G, Romano AD, Cassano T, Bellanti F, Altomare E, Vendemiale G. Principles and therapeutic relevance for targeting mitochondria in aging and neurodegenerative diseases. Current Pharmaceutical Design. 2011;17(20):2036–2055. doi: 10.2174/138161211796904740. [DOI] [PubMed] [Google Scholar]
- 113.Widlansky ME, Gutterman DD. Regulation of endothelial function by mitochondrial reactive oxygen species. Antioxidants and Redox Signaling. 2011;15(6):1517–1530. doi: 10.1089/ars.2010.3642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Chen S-D, Yang D-I, Lin T-K, Shaw F-Z, Liou C-W, Chuang Y-C. Roles of oxidative stress, apoptosis, PGC-1 and mitochondrial biogenesis in cerebral ischemia. International Journal of Molecular Sciences. 2011;12(10):7199–7215. doi: 10.3390/ijms12107199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Hensley K, Maidt ML, Yu Z, Sang H, Markesbery WR, Floyd RA. Electrochemical analysis of protein nitrotyrosine and dityrosine in the Alzheimer brain indicates region-specific accumulation. Journal of Neuroscience. 1998;18(20):8126–8132. doi: 10.1523/JNEUROSCI.18-20-08126.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Allan Butterfield D, Castegna A, Lauderback CM, Drake J. Evidence that amyloid beta-peptide-induced lipid peroxidation and its sequelae in Alzheimer’s disease brain contribute to neuronal death. Neurobiology of Aging. 2002;23(5):655–664. doi: 10.1016/s0197-4580(01)00340-2. [DOI] [PubMed] [Google Scholar]
- 117.Dexter DT, Carter CJ, Wells FR, et al. Basal lipid peroxidation in substantia nigra is increased in Parkinson’s disease. Journal of Neurochemistry. 1989;52(2):381–389. doi: 10.1111/j.1471-4159.1989.tb09133.x. [DOI] [PubMed] [Google Scholar]
- 118.Pedersen WA, Fu W, Keller JN, et al. Protein modification by the lipid peroxidation product 4-hydroxynonenal in the spinal cords of amyotrophic lateral sclerosis patients. Annals of Neurology. 1998;44(5):819–824. doi: 10.1002/ana.410440518. [DOI] [PubMed] [Google Scholar]
- 119.Pearce RKB, Owen A, Daniel S, Jenner P, Marsden CD. Alterations in the distribution of glutathione in the substantia nigra in Parkinson’s disease. Journal of Neural Transmission. 1997;104(6-7):661–677. doi: 10.1007/BF01291884. [DOI] [PubMed] [Google Scholar]
- 120.Sofic E, Paulus W, Jellinger K, Riederer P, Youdim MBH. Selective increase of iron in substantia nigra zona compacta of Parkinsonian brains. Journal of Neurochemistry. 1991;56(3):978–982. doi: 10.1111/j.1471-4159.1991.tb02017.x. [DOI] [PubMed] [Google Scholar]
- 121.Jellinger KA, Kienzl E, Rumpelmaier G, et al. Iron and ferritin in substantia nigra in Parkinson’s disease. Advances in neurology. 1993;60:267–272. [PubMed] [Google Scholar]
- 122.Zemlan FP, Thienhaus OJ, Bosmann HB. Superoxide dismutase activity in Alzheimer’s disease: possible mechanism for paired helical filament formation. Brain Research. 1989;476(1):160–162. doi: 10.1016/0006-8993(89)91550-3. [DOI] [PubMed] [Google Scholar]
- 123.Pappolla MA, Omar RA, Kim KS, Robakis NK. Immunohistochemical evidence of antioxidant stress in Alzheimer's disease. American Journal of Pathology. 1992;140(3):621–628. [PMC free article] [PubMed] [Google Scholar]
- 124.Perry TL, Godin DV, Hansen S. Parkinson’s disease: a disorder due to nigral glutathione deficiency? Neuroscience Letters. 1982;33(3):305–310. doi: 10.1016/0304-3940(82)90390-1. [DOI] [PubMed] [Google Scholar]
- 125.Warita H, Hayashi T, Murakami T, Manabe Y, Abe K. Oxidative damage to mitochondrial DNA in spinal motoneurons of transgenic ALS mice. Molecular Brain Research. 2001;89(1-2):147–152. doi: 10.1016/s0169-328x(01)00029-8. [DOI] [PubMed] [Google Scholar]
- 126.Zhang J, Graham DG, Montine TJ, Ho YS. Enhanced N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine toxicity in mice deficient in CuZn-superoxide dismutase or glutathione peroxidase. Journal of Neuropathology and Experimental Neurology. 2000;59(1):53–61. doi: 10.1093/jnen/59.1.53. [DOI] [PubMed] [Google Scholar]
- 127.Resende R, Moreira PI, Proença T, et al. Brain oxidative stress in a triple-transgenic mouse model of Alzheimer disease. Free Radical Biology and Medicine. 2008;44(12):2051–2057. doi: 10.1016/j.freeradbiomed.2008.03.012. [DOI] [PubMed] [Google Scholar]
- 128.Jagatha B, Mythri RB, Vali S, Bharath MMS. Curcumin treatment alleviates the effects of glutathione depletion in vitro and in vivo: therapeutic implications for Parkinson’s disease explained via in silico studies. Free Radical Biology and Medicine. 2008;44(5):907–917. doi: 10.1016/j.freeradbiomed.2007.11.011. [DOI] [PubMed] [Google Scholar]
- 129.Chen J, Tang XQ, Zhi JL, et al. Curcumin protects PC12 cells against 1-methyl-4-phenylpyridinium ion-induced apoptosis by bcl-2-mitochondria-ROS-iNOS pathway. Apoptosis. 2006;11(6):943–953. doi: 10.1007/s10495-006-6715-5. [DOI] [PubMed] [Google Scholar]
- 130.Mishra S, Mishra M, Seth P, Kumar Sharma S. Tetrahydrocurcumin confers protection against amyloid β-induced toxicity. NeuroReport. 2011;22(1):23–27. doi: 10.1097/WNR.0b013e328341e141. [DOI] [PubMed] [Google Scholar]
- 131.Mythri RB, Harish G, Dubey SK, Misra K, Srinivas Bharath MM. Glutamoyl diester of the dietary polyphenol curcumin offers improved protection against peroxynitrite-mediated nitrosative stress and damage of brain mitochondria in vitro: implications for Parkinson’s disease. Molecular and Cellular Biochemistry. 2011;347(1-2):135–143. doi: 10.1007/s11010-010-0621-4. [DOI] [PubMed] [Google Scholar]
- 132.Srividhya R, Zarkovic K, Stroser M, Waeg G, Zarkovic N, Kalaiselvi P. Mitochondrial alterations in aging rat brain: effective role of (-)-epigallo catechin gallate. International Journal of Developmental Neuroscience. 2009;27(3):223–231. doi: 10.1016/j.ijdevneu.2009.01.003. [DOI] [PubMed] [Google Scholar]
- 133.Levites Y, Weinreb O, Maor G, Youdim MBH, Mandel S. Green tea polyphenol (-)-epigallocatechin-3-gallate prevents N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced dopaminergic neurodegeneration. Journal of Neurochemistry. 2001;78(5):1073–1082. doi: 10.1046/j.1471-4159.2001.00490.x. [DOI] [PubMed] [Google Scholar]
- 134.Na H-K, Surh Y-J. Modulation of Nrf2-mediated antioxidant and detoxifying enzyme induction by the green tea polyphenol EGCG. Food and Chemical Toxicology. 2008;46(4):1271–1278. doi: 10.1016/j.fct.2007.10.006. [DOI] [PubMed] [Google Scholar]
- 135.Murphy MP. Selective targeting of bioactive compounds to mitochondria. Trends in Biotechnology. 1997;15(8):326–330. doi: 10.1016/S0167-7799(97)01068-8. [DOI] [PubMed] [Google Scholar]
- 136.Yamada Y, Akita H, Kogure K, Kamiya H, Harashima H. Mitochondrial drug delivery and mitochondrial disease therapy—an approach to liposome-based delivery targeted to mitochondria. Mitochondrion. 2007;7(1-2):63–71. doi: 10.1016/j.mito.2006.12.003. [DOI] [PubMed] [Google Scholar]
- 137.Yousif LF, Stewart KM, Kelley SO. Targeting mitochondria with organelle-specific compounds: strategies and applications. ChemBioChem. 2009;10(12):1939–1950. doi: 10.1002/cbic.200900185. [DOI] [PubMed] [Google Scholar]
- 138.Yousif LF, Stewart KM, Horton KL, Kelley SO. Mitochondria-penetrating peptides: sequence effects and model cargo transport. ChemBioChem. 2009;10(12):2081–2088. doi: 10.1002/cbic.200900017. [DOI] [PubMed] [Google Scholar]
- 139.Jain SK, Levine SN, Duett J, Hollier B. Reduced vitamin E and increased lipofuscin products in erythrocytes of diabetic rats. Diabetes. 1991;40(10):1241–1244. doi: 10.2337/diab.40.10.1241. [DOI] [PubMed] [Google Scholar]
- 140.Jain SK, McVie R. Effect of glycemic control, race (white versus black), and duration of diabetes on reduced glutathione content in erythrocytes of diabetic patients. Metabolism: Clinical and Experimental. 1994;43(3):306–309. doi: 10.1016/0026-0495(94)90097-3. [DOI] [PubMed] [Google Scholar]
- 141.Hernandez-Mijares A, Rocha M, Apostolova N, et al. Mitochondrial complex i impairment in leukocytes from type 2 diabetic patients. Free Radical Biology and Medicine. 2011;50(10):1215–1221. doi: 10.1016/j.freeradbiomed.2011.01.019. [DOI] [PubMed] [Google Scholar]
- 142.Brownlee M. Biochemistry and molecular cell biology of diabetic complications. Nature. 2001;414(6865):813–820. doi: 10.1038/414813a. [DOI] [PubMed] [Google Scholar]
- 143.Ahmad FK, He Z, King GL. Molecular targets of diabetic cardiovascular complications. Current Drug Targets. 2005;6(4):487–494. doi: 10.2174/1389450054021990. [DOI] [PubMed] [Google Scholar]
- 144.Du X, Edelstein D, Obici S, Higham N, Zou MH, Brownlee M. Insulin resistance reduces arterial prostacyclin synthase and eNOS activities by increasing endothelial fatty acid oxidation. Journal of Clinical Investigation. 2006;116(4):1071–1080. doi: 10.1172/JCI23354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Martín-Hernández E, García-Silva MT, Vara J, et al. Renal pathology in children with mitochondrial diseases. Pediatric Nephrology. 2005;20(9):1299–1305. doi: 10.1007/s00467-005-1948-z. [DOI] [PubMed] [Google Scholar]
- 146.Krauss S, Zhang CY, Scorrano L, et al. Superoxide-mediated activation of uncoupling protein 2 causes pancreatic β cell dysfunction. Journal of Clinical Investigation. 2003;112(12):1831–1842. doi: 10.1172/JCI19774. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Ceriello A, Morocutti A, Mercuri F, et al. Defective intracellular antioxidant enzyme production in type 1 diabetic patients with nephropathy. Diabetes. 2000;49(12):2170–2177. doi: 10.2337/diabetes.49.12.2170. [DOI] [PubMed] [Google Scholar]
- 148.Möllsten A, Marklund SL, Wessman M, et al. A functional polymorphism in the manganese superoxide dismutase gene and diabetic nephropathy. Diabetes. 2007;56(1):265–269. doi: 10.2337/db06-0698. [DOI] [PubMed] [Google Scholar]
- 149.Brezniceanu ML, Liu F, Wei CC, et al. Catalase overexpression attenuates angiotensinogen expression and apoptosis in diabetic mice. Kidney International. 2007;71(9):912–923. doi: 10.1038/sj.ki.5002188. [DOI] [PubMed] [Google Scholar]
- 150.Haffner SM, Lehto S, Rönnemaa T, Pyörälä K, Laakso M. Mortality from coronary heart disease in subjects with type 2 diabetes and in nondiabetic subjects with and without prior myocardial infarction. The New England Journal of Medicine. 1998;339(4):229–234. doi: 10.1056/NEJM199807233390404. [DOI] [PubMed] [Google Scholar]
- 151.Semenkovich CF. Insulin resistance and atherosclerosis. Journal of Clinical Investigation. 2006;116(7):1813–1822. doi: 10.1172/JCI29024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Nishikawa T, Edelstein D, Du XL, et al. Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature. 2000;404(6779):787–790. doi: 10.1038/35008121. [DOI] [PubMed] [Google Scholar]
- 153.Ceriello A, Testa R. Antioxidant anti-inflammatory treatment in type 2 diabetes. Diabetes Care. 2009;32:S232–236. doi: 10.2337/dc09-S316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Blum S, Vardi M, Brown JB, et al. Vitamin e reduces cardiovascular disease in individuals with diabetes mellitus and the haptoglobin 2-2 genotype. Pharmacogenomics. 2010;11(5):675–684. doi: 10.2217/pgs.10.17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Milman U, Blum S, Shapira C, et al. Vitamin E supplementation reduces cardiovascular events in a subgroup of middle-aged individuals with both type 2 diabetes mellitus and the haptoglobin 2-2 genotype: a prospective double-blinded clinical trial. Arteriosclerosis, Thrombosis, and Vascular Biology. 2008;28(2):341–347. doi: 10.1161/ATVBAHA.107.153965. [DOI] [PubMed] [Google Scholar]
- 156.Green K, Brand MD, Murphy MP. Prevention of Mitochondrial Oxidative Damage As A Therapeutic Strategy in Diabetes. Diabetes. 2004;53(1):S110–S118. doi: 10.2337/diabetes.53.2007.s110. [DOI] [PubMed] [Google Scholar]
- 157.Chacko BK, Reily C, Srivastava A, et al. Prevention of diabetic nephropathy in Ins2+/-AkitaJ mice by the mitochondria-targeted therapy MitoQ. Biochemical Journal. 2010;432(1):9–19. doi: 10.1042/BJ20100308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Green DR, Reed JC. Mitochondria and apoptosis. Science. 1998;281(5381):1309–1312. doi: 10.1126/science.281.5381.1309. [DOI] [PubMed] [Google Scholar]
- 159.Debatin KM, Poncet D, Kroemer G. Chemotherapy: targeting the mitochondrial cell death pathway. Oncogene. 2002;21(57):8786–8803. doi: 10.1038/sj.onc.1206039. [DOI] [PubMed] [Google Scholar]
- 160.Armstrong JS. Mitochondria: a target for cancer therapy. British Journal of Pharmacology. 2006;147(3):239–248. doi: 10.1038/sj.bjp.0706556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Gogvadze V, Orrenius S, Zhivotovsky B. Mitochondria in cancer cells: what is so special about them? Trends in Cell Biology. 2008;18(4):165–173. doi: 10.1016/j.tcb.2008.01.006. [DOI] [PubMed] [Google Scholar]
- 162.Radad K, Rausch WD, Gille G. Rotenone induces cell death in primary dopaminergic culture by increasing ROS production and inhibiting mitochondrial respiration. Neurochemistry International. 2006;49(4):379–386. doi: 10.1016/j.neuint.2006.02.003. [DOI] [PubMed] [Google Scholar]
- 163.Lindahl PE, Öberg KE. The effect of rotenone on respiration and its point of attack. Experimental Cell Research. 1961;23(2):228–237. doi: 10.1016/0014-4827(61)90033-7. [DOI] [PubMed] [Google Scholar]
- 164.Wolvetang EJ, Johnson KL, Krauer K, Ralph SJ, Linnane AW. Mitochondrial respiratory chain inhibitors induce apoptosis. FEBS Letters. 1994;339(1-2):40–44. doi: 10.1016/0014-5793(94)80380-3. [DOI] [PubMed] [Google Scholar]
- 165.Isenberg JS, Klaunig JE. Role of the mitochondrial membrane permeability transition (MPT) in rotenone-induced apoptosis in liver cells. Toxicological Sciences. 2000;53(2):340–351. doi: 10.1093/toxsci/53.2.340. [DOI] [PubMed] [Google Scholar]
- 166.Isenberg JS, Kolaja KL, Ayoubi SA, Watkins JB, Klaunig JE. Inhibition of WY-14,643 induced hepatic lesion growth in mice by rotenone. Carcinogenesis. 1997;18(8):1511–1519. doi: 10.1093/carcin/18.8.1511. [DOI] [PubMed] [Google Scholar]
- 167.Armstrong JS, Hornung B, Lecane P, Jones DP, Knox SJ. Rotenone-induced G2/M cell cycle arrest and apoptosis in a human B lymphoma cell line PW. Biochemical and Biophysical Research Communications. 2001;289(5):973–978. doi: 10.1006/bbrc.2001.6054. [DOI] [PubMed] [Google Scholar]
- 168.Deng YT, Huang HC, Lin JK. Rotenone induces apoptosis in MCF-7 human breast cancer cell-mediated ROS through JNK and p38 signaling. Molecular Carcinogenesis. 2010;49(2):141–151. doi: 10.1002/mc.20583. [DOI] [PubMed] [Google Scholar]
- 169.Chung WG, Miranda CL, Maier CS. Epigallocatechin gallate (EGCG) potentiates the cytotoxicity of rotenone in neuroblastoma SH-SY5Y cells. Brain Research. 2007;1176(1):133–142. doi: 10.1016/j.brainres.2007.07.083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Tada-Oikawa S, Hiraku Y, Kawanishi M, Kawanishi S. Mechanism for generation of hydrogen peroxide and change of mitochondrial membrane potential during rotenone-induced apoptosis. Life Sciences. 2003;73(25):3277–3288. doi: 10.1016/j.lfs.2003.06.013. [DOI] [PubMed] [Google Scholar]
- 171.Moreira PI, Custódio J, Moreno A, Oliveira CR, Santos MS. Tamoxifen and estradiol interact with the flavin mononucleotide site of complex I leading to mitochondrial failure. Journal of Biological Chemistry. 2006;281(15):10143–10152. doi: 10.1074/jbc.M510249200. [DOI] [PubMed] [Google Scholar]
- 172.Kallio A, Zheng A, Dahllund J, Heiskanen KM, Härkönen P. Role of mitochondria in tamoxifen-induced rapid death of MCF-7 breast cancer cells. Apoptosis. 2005;10(6):1395–1410. doi: 10.1007/s10495-005-2137-z. [DOI] [PubMed] [Google Scholar]
- 173.Alston TA, Mela L, Bright HJ. 3-Nitropropionate, the toxic substance of Indigofera, is a suicide inactivator of succinate dehydrogenase. Proceedings of the National Academy of Sciences of the United States of America. 1977;74(9):3767–3771. doi: 10.1073/pnas.74.9.3767. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Coles CJ, Edmondson DE, Singer TP. Inactivation of succinate dehydrogenase by 3-nitropropionate. Journal of Biological Chemistry. 1979;254(12):5161–5167. [PubMed] [Google Scholar]
- 175.Huang LS, Sun G, Cobessi D, et al. 3-Nitropropionic acid is a suicide inhibitor of mitochondrial respiration that, upon oxidation by complex II, forms a covalent adduct with a catalytic base arginine in the active site of the enzyme. Journal of Biological Chemistry. 2006;281(9):5965–5972. doi: 10.1074/jbc.M511270200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Bacsi A, Woodberry M, Widger W, et al. Localization of superoxide anion production to mitochondrial electron transport chain in 3-NPA-treated cells. Mitochondrion. 2006;6(5):235–244. doi: 10.1016/j.mito.2006.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Wang J, Green PS, Simpkins JW. Estradiol protects against ATP depletion, mitochondrial membrane potential decline and the generation of reactive oxygen species induced by 3-nitroproprionic acid in SK-N-SH human neuroblastoma cells. Journal of Neurochemistry. 2001;77(3):804–811. doi: 10.1046/j.1471-4159.2001.00271.x. [DOI] [PubMed] [Google Scholar]
- 178.Dong LF, Low P, Dyason JC, et al. α-Tocopheryl succinate induces apoptosis by targeting ubiquinone-binding sites in mitochondrial respiratory complex II. Oncogene. 2008;27(31):4324–4335. doi: 10.1038/onc.2008.69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Neuzil J, Weber T, Schröder A, et al. Induction of cancer cell apoptosis by α-tocopheryl succinate: molecular pathways and structural requirements. The FASEB Journal. 2001;15(2):403–415. doi: 10.1096/fj.00-0251com. [DOI] [PubMed] [Google Scholar]
- 180.Wu K, Zhao Y, Li GC, Yu WP. c-Jun N-terminal kinase is required for vitamin E succinate-induced apoptosis in human gastric cancer cells. World Journal of Gastroenterology. 2004;10(8):1110–1114. doi: 10.3748/wjg.v10.i8.1110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Wu K, Zhao Y, Liu BH, et al. RRR-α-tocopheryl succinate inhibits human gastric cancer SGC-7901 cell growth by inducing apoptosis and DNA synthesis arrest. World Journal of Gastroenterology. 2002;8(1):26–30. doi: 10.3748/wjg.v8.i1.26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Yu W, Qiao Yin Liao, Hantash FM, Sanders BG, Kline K. Activation of extracellular signal-regulated kinase and c-Jun-NH2-terminal kinase but not p38 mitogen-activated protein kinases is required for RRR-α-tocopheryl succinate-induced apoptosis of human breast cancer cells. Cancer Research. 2001;61(17):6569–6576. [PubMed] [Google Scholar]
- 183.Weber T, Lu M, Andera L, et al. Vitamin E succinate is a potent novel antineoplastic agent with high selectivity and cooperativity with tumor necrosis factorrelated apoptosis-inducing ligand (Apo2 ligand) in vivo. Clinical Cancer Research. 2002;8(3):863–869. [PubMed] [Google Scholar]
- 184.Kline K, Yu W, Sanders BG. Vitamin E and breast cancer. Journal of Nutrition. 2004;134(12, supplement):3458S–3462S. doi: 10.1093/jn/134.12.3458S. [DOI] [PubMed] [Google Scholar]
- 185.Yu W, Sanders BG, Kline K. RRR-α-tocopheryl succinate-induced apoptosis of human breast cancer cells involves bax translocation to mitochondrial. Cancer Research. 2003;63(10):2483–2491. [PubMed] [Google Scholar]
- 186.Wang XF, Dong L, Zhao Y, Tomasetti M, Wu K, Neuzil J. Vitamin E analogues as anticancer agents: lessons from studies with α-tocopheryl succinate. Molecular Nutrition and Food Research. 2006;50(8):675–685. doi: 10.1002/mnfr.200500267. [DOI] [PubMed] [Google Scholar]
- 187.Nakayama K, Okamoto F, Harada Y. Antimycin A: isolation from a new Streptomyces and activity against rice plant blast fungi. The Journal of Antibiotics. 1956;9(2):63–66. [PubMed] [Google Scholar]
- 188.Campo ML, Kinnally KW, Tedeschi H. The effect of antimycin A on mouse liver inner mitochondrial membrane channel activity. Journal of Biological Chemistry. 1992;267(12):8123–8127. [PubMed] [Google Scholar]
- 189.Alexandre A, Lehninger AL. Bypasses of the antimycin A block of mitochondrial electron transport in relation to ubisemiquinone function. Biochimica et Biophysica Acta. 1984;767(1):120–129. doi: 10.1016/0005-2728(84)90086-0. [DOI] [PubMed] [Google Scholar]
- 190.Balaban RS, Nemoto S, Finkel T. Mitochondria, oxidants, and aging. Cell. 2005;120(4):483–495. doi: 10.1016/j.cell.2005.02.001. [DOI] [PubMed] [Google Scholar]
- 191.You KR, Wen J, Lee ST, Kim DG. Cytochrome c oxidase subunit III: a molecular marker for N-(4-hydroxyphenyl)retinamide-induced oxidative stress in hepatoma cells. Journal of Biological Chemistry. 2002;277(6):3870–3877. doi: 10.1074/jbc.M109284200. [DOI] [PubMed] [Google Scholar]
- 192.Wu JM, DiPietrantonio AM, Hsieh TC. Mechanism of fenretinide (4-HPR)-induced cell death. Apoptosis. 2001;6(5):377–388. doi: 10.1023/a:1011342220621. [DOI] [PubMed] [Google Scholar]
- 193.Li X, Ling W, Pennisi A, Khan S, Yaccoby S. Fenretinide inhibits myeloma cell growth, osteoclastogenesis and osteoclast viability. Cancer Letters. 2009;284(2):175–181. doi: 10.1016/j.canlet.2009.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Suzuki S, Higuchi M, Proske RJ, Oridate N, Hong WK, Lotan R. Implication of mitochondria-derived reactive oxygen species, cytochrome C and caspase-3 in N-(4-Hydroxyphenyl)retinamide-induced apoptosis in cervical carcinoma cells. Oncogene. 1999;18(46):6380–6387. doi: 10.1038/sj.onc.1203024. [DOI] [PubMed] [Google Scholar]
- 195.Kim HJ, Chakravarti N, Oridate N, Choe C, Claret FX, Lotan R. N-(4-hydroxyphenyl)retinamide-induced apoptosis triggered by reactive oxygen species is mediated by activation of MAPKs in head and neck squamous carcinoma cells. Oncogene. 2006;25(19):2785–2794. doi: 10.1038/sj.onc.1209303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Hail Jr. N, Kim HJ, Lotan R. Mechanisms of fenretinide-induced apoptosis. Apoptosis. 2006;11(10):1677–1694. doi: 10.1007/s10495-006-9289-3. [DOI] [PubMed] [Google Scholar]
- 197.Formelli F, Barua AB, Olson JA. Bioactivities of N-(4-hydroxyphenyl) retinamide and retinoyl β-glucuronide. The FASEB Journal. 1996;10(9):1014–1024. doi: 10.1096/fasebj.10.9.8801162. [DOI] [PubMed] [Google Scholar]
- 198.Valero JG, Sancey L, Kucharczak J, et al. Bax-derived membrane-active peptides act as potent and direct inducers of apoptosis in cancer cells. Journal of Cell Science. 2011;124(4):556–564. doi: 10.1242/jcs.076745. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Venditto VJ, Simanek EE. Cancer therapies utilizing the camptothecins: a review of the in vivo literature. Molecular Pharmaceutics. 2010;7(2):307–349. doi: 10.1021/mp900243b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Hande KR. Etoposide: four decades of development of a topoisomerase II inhibitor. European Journal of Cancer. 1998;34(10):1514–1521. doi: 10.1016/s0959-8049(98)00228-7. [DOI] [PubMed] [Google Scholar]
- 201.Jordan MA, Wilson L. Microtubules as a target for anticancer drugs. Nature Reviews Cancer. 2004;4(4):253–265. doi: 10.1038/nrc1317. [DOI] [PubMed] [Google Scholar]
- 202.Douer D, Tollman MS. Arsenic trioxide: new clinical experience with an old medication in hematologic malignancies. Journal of Clinical Oncology. 2005;23(10):2396–2410. doi: 10.1200/JCO.2005.10.217. [DOI] [PubMed] [Google Scholar]
- 203.Petronellia A, Pannitterib G, Testaa U. Triterpenoids as new promising anticancer drugs. Anti-Cancer Drugs. 2009;20(10):880–892. doi: 10.1097/CAD.0b013e328330fd90. [DOI] [PubMed] [Google Scholar]
- 204.Shehzad A, Wahid F, Lee YS. Curcumin in cancer chemoprevention: molecular targets, pharmacokinetics, bioavailability, and clinical trials. Archiv der Pharmazie. 2010;343(9):489–499. doi: 10.1002/ardp.200900319. [DOI] [PubMed] [Google Scholar]
- 205.Kolesar J, Brundage RC, Pomplun M, et al. Population pharmacokinetics of 3-aminopyridine-2-carboxaldehyde thiosemicarbazone (Triapine®) in cancer patients. Cancer Chemotherapy and Pharmacology. 2010:1–8. doi: 10.1007/s00280-010-1331-z. [DOI] [PMC free article] [PubMed] [Google Scholar]