Reviews

Paradoxical Roles of Antioxidant Enzymes: Basic Mechanisms and Health Implications

Published Online:https://doi.org/10.1152/physrev.00010.2014

Abstract

Reactive oxygen species (ROS) and reactive nitrogen species (RNS) are generated from aerobic metabolism, as a result of accidental electron leakage as well as regulated enzymatic processes. Because ROS/RNS can induce oxidative injury and act in redox signaling, enzymes metabolizing them will inherently promote either health or disease, depending on the physiological context. It is thus misleading to consider conventionally called antioxidant enzymes to be largely, if not exclusively, health protective. Because such a notion is nonetheless common, we herein attempt to rationalize why this simplistic view should be avoided. First we give an updated summary of physiological phenotypes triggered in mouse models of overexpression or knockout of major antioxidant enzymes. Subsequently, we focus on a series of striking cases that demonstrate “paradoxical” outcomes, i.e., increased fitness upon deletion of antioxidant enzymes or disease triggered by their overexpression. We elaborate mechanisms by which these phenotypes are mediated via chemical, biological, and metabolic interactions of the antioxidant enzymes with their substrates, downstream events, and cellular context. Furthermore, we propose that novel treatments of antioxidant enzyme-related human diseases may be enabled by deliberate targeting of dual roles of the pertaining enzymes. We also discuss the potential of “antioxidant” nutrients and phytochemicals, via regulating the expression or function of antioxidant enzymes, in preventing, treating, or aggravating chronic diseases. We conclude that “paradoxical” roles of antioxidant enzymes in physiology, health, and disease derive from sophisticated molecular mechanisms of redox biology and metabolic homeostasis. Simply viewing antioxidant enzymes as always being beneficial is not only conceptually misleading but also clinically hazardous if such notions underpin medical treatment protocols based on modulation of redox pathways.

I. INTRODUCTION

Antioxidant enzymes are often discussed in scientific research and daily life as key players of metabolism that promote healthy cells, tissues, and organisms. The term best relates to enzymes that lower the levels of reactive oxygen species (ROS) and reactive nitrogen species (RNS), or counteract their downstream cellular effects of excessive oxidation. ROS and RNS are produced from aerobic biogenesis or by oxidative enzymes, and being chemically reactive they have the capacity to damage cellular components. Nature has evolved three layers of antioxidant defense in the body. Small molecular antioxidants, including uric acid, glutathione (GSH), and vitamins C and E, offer the first line of defense to scavenge ROS/RNS directly and thus prevent or delay the initiation of various oxidative stresses. Damage-removing or repairing enzymes function as the last defense to regenerate biomolecules damaged from oxidative injury. Between these two layers, antioxidant enzymes serve as an intermediate defense to detoxify ROS/RNS into less reactive species. Superoxide (O2) and hydrogen peroxide (H2O2) represent arguably the best-known and most-produced ROS, with the former scavenged by superoxide dismutases (SOD) (69) and the latter by catalase (CAT) (337), glutathione peroxidases (GPX) (62), and peroxiredoxins (PRX) (562).1 Thioredoxin reductases (TrxR) are in addition required to maintain functions of thioredoxins (Trx), PRX, methionine sulfoxide reductases (Msr) (499) and many other redox-regulatory enzymes/proteins by regenerating protein thiols (17, 411) in parallel with glutaredoxins (Grx) utilizing GSH to catalyze reduction of protein disulfide substrates (180, 395). These ROS-metabolizing and reductive enzymes, which also play important roles in RNS homeostasis, are widely considered to be the major antioxidant enzymes and are the focus of this review article.

During the past two decades, developments of antioxidant enzyme gene knockout and overexpression mouse models (Table 1) have enabled not only verifying the “anticipated” metabolic health-promoting functions of these enzymes, but also revealing the often neglected “paradoxical” roles of antioxidant enzymes triggering metabolic disorders. It has indeed become clear that many antioxidant enzymes are more than just protective ROS/RNS scavengers. They regulate many redox signaling pathways and may also exhibit pro-oxidant functions or functions independent of their redox activities.

Table 1. Commonly used mouse models for antioxidant enzyme overexpression and knockout

Overexpression
Knockout
Gene Product Nature of the Transgene Altered Site Reference Nos. Disrupted Gene Altered Site Reference Nos.
Cu,Zn-superoxide dismutase (SOD1) The entire human SOD1 gene contained in a 14.5-kb genomic fragment Brain, liver, heart, and lung 87, 170, 681 Sod1 Global 264, 285, 443, 556
The entire human SOD1 gene contained in a 64-kb genomic fragment Brain, heart, kidney, liver, lung, skeletal muscle, and spleen 103
Mn-superoxide dismutase (SOD2) A human SOD2 expression construct driven by 3.7 kb of the promoter and 5′ flanking sequences of the human surfactant protein C gene Lung 703 Sod2 Global and tissue specific 291, 370, 392, 563, 621
A human SOD2 expression construct controlled by 3 kb of 5′ flanking sequence plus 5′ untranslated region and intron 1 of the human β-actin gene Brain, eye, heart, lung, skeletal muscle, spleen, and tongue 266, 494, 727
The entire mouse Sod2 gene contained in a 14-kb genomic fragment Brain, heart, kidney, liver, and lung 542
A human SOD2 expression construct driven by 570 bp of 5′ flanking sequence and promoter of the rat insulin I gene Pancreatic β-cells 99
A human SOD2 expression vector controlled by a 2-kb promoter and 10-kb enhancer of the mouse Tie2 (a vascular endothelial-specific receptor tyrosine kinase) gene Endothelial cells 226
A human SOD2 expression construct controlled by a 5.5-kb mouse genomic fragment containing the last intron of the β-myosin heavy chain (MHC) gene to exon 3 of the α-MHC gene Heart 597
Extracellular superoxide dismutase (SOD3) A human SOD3 expression construct controlled by 3 kb of 5′ flanking sequence plus 5′ untranslated region and intron 1 of the human β-actin gene Brain, heart, and skeletal muscle 509 Sod3 Global 81
A human SOD3 expression construct driven by 3.7 kb of the promoter and 5′ flanking sequences of the human surfactant protein C gene Lung 189
Catalase (CAT) A rat CAT expression construct downstream of a 5.5-kb mouse genomic fragment containing the last intron of the β-myosin heavy chain (MHC) gene to exon 3 of the α-MHC gene Heart 319 Cat Global 268
A human CAT expression construct controlled by a 2.8-kb mouse α-fetoprotein enhancer element I fused to 1.8 kb of the human β-globin promoter Liver and gut 486
A rat CAT expression construct driven by 570 bp of 5′ flanking sequence and promoter of the rat insulin I gene Pancreatic β-cells 717
Three human CAT expression constructs (peroxisome-, nucleus-, and mitochondria-targeted) driven by the cytomegalovirus enhancer element and a chicken β-actin promoter Brain, heart, kidney, skeletal muscle, and spleen 584, 585
The entire human CAT gene contained in a 80-kb genomic fragment Brain, heart, kidney, liver, lung, skeletal muscle, and spleen 103
Glutathione peroxidase 1 (GPX1) A human GPX1 expression construct controlled by the promoter, exon 1, and intron 1 of the mouse hydroxymethylglutaryl-coenzyme A reductase gene Brain, heart, kidney, and liver 457 Gpx1 Global 141, 172, 265
A human GPX1 expression construct controlled by rat insulin II promoter Pancreas 244
The entire mouse Gpx1 gene contained in a 5.3-kb genomic fragment Brain, eye, heart, lung, skeletal muscle, spleen, pancreas, and tongue 109, 716, 733
Gastrointestinal glutathione peroxidase (GPX2) Gpx2 Global 176, 186
Glutathione peroxidase 3 (GPX3) A human GPX3 expression construct controlled by the promoter, exon 1, and intron 1 of the mouse hydroxymethylglutaryl-coenzyme A reductase gene Kidney, brain, and lung 457 Gpx3 Global 311
Phospholipid hydroperoxide glutathione peroxidase (GPX4) A rat mitochondrion-targeted Gpx4 expression driven by the human cytomegalovirus immediate early enhancer and chicken β-actin promoter Mitochondria of the heart 136 Gpx4 Global, neurons, spermatocytes, cytosol, mitochondria, nucleus 52, 292, 293, 581, 590, 725
A 50-kb genomic clone containing the entire human GPX4 gene Cerebral cortex, heart, skeletal muscle, kidney, liver, and testes 548
Peroxiredoxin I (PRX1) Prx1 Global 338, 484
Peroxiredoxin II (PRX2) Prx2 Global 375
Peroxiredoxin III (PRX3) A rat Prx3 expression construct driven by the cytomegalovirus promoter Mitochondria of the heart 440 Prx3 Global 389
Peroxiredoxin IV (PRX4) A human PRX4 expression construct driven by the enhancer and promoter of the human cytomegalovirus immediate early gene Brain, heart, kidney, pancreas, and testis 155 Prx4 Global 300
Peroxiredoxin VI (PRX6) A 16.8-kb genomic fragment containing the entire mouse Prx6 gene isolated from 129SvJ mice Intestine, kidney, liver, lung, and epithelial cells of all tissues 531 Prx6 Global 461, 683
Thioredoxin 1 (TXN1 or Trx1) A human TXN1 expression construct controlled by the human insulin promoter and exons and introns of the rabbit β globin gene Pancreas 281 Txn1/Trx1 Global 437
A human TXN1 expression construct driven by 3 kb of 5′ flanking sequence plus 5′ untranslated region and intron 1 of the human β-actin gene. Brain, heart, kidney, liver, lung, skeletal muscle, spleen, and tongue 3, 661
The structure of the human TXN1 transgene is not described. Brain, heart, kidney, liver, lung, and skin 636
Thioredoxin 2 (TXN2 or Trx2) The structure of the human TXN2 transgene is not described Heart 696 Txn2/Trx2 Global 490
Thioredoxin reductase 1 (TrxR1) Txnrd1 Global, neurons, liver, heart 51, 79, 302,305, 617
Thioredoxin reductase 2 (TrxR2) Txnrd2 Global, neurons, heart 123, 617
Selenoprotein P (SEPP1) Sepp1 Global 259, 583
Glutaredoxin 1 (Grx1or Glrx1) A human Grx1 expresssion construct driven by the human beta-actin promoter Endothelia, heart, muscle 478 Glrx/Grx1 Global 4, 267, 270
Glutaredoxin 2 (Grx2 or Glrx2) Grx2 Global 426, 427, 710
Selenocysteine tRNA (Trsp) A 1.93-kb genomic DNA containing the entire mouse Trsp gene with mutations of TC at position 9 and AG at position 37 Brain, kidney, liver, and testes 471 Trsp Global and several tissue-specific knockouts 54, 356, 361
Methionine sulfoxide reductase A (MSRA) Three mouse MsrA expression constructs (wild-type, mitochondria-targeted, and cytosolic) controlled by the cytomegalovirus enhancer and chicken β-actin promoter Liver, skin fibroblasts 752 MsrA Global 468
Methionine sulfoxide reductase B (MSRB) MsrB1 Global 190

A number of chronic diseases are associated with genetic or metabolic alterations of antioxidant enzymes, displaying either lower or increased activities, depending on the actual enzyme and disease. Meanwhile, certain “antioxidant” nutrients and phytochemicals are able to regulate antioxidant enzyme expressions or functions with health implications. Therefore, it is important to have a better understanding of the “paradoxical” functions of antioxidant enzymes in physiology, which explain how their overexpression can promote disease or their deletion can be health-promoting. It is our goal that this review will help to support awareness of the involved molecular mechanisms and thus be useful in advancing a more balanced view of antioxidant enzymes and redox biology in medicine.

II. IMPACTS OF KNOCKOUT AND OVEREXPRESSION

Superoxide- and H2O2-metabolizing enzymes, including SOD, catalase, GPX, and PRX, are generally considered to be the primary antioxidant enzyme defense system in the body. However, the only antioxidant enzymes that have thus far been found to be essential for mouse embryonic development and thus lethal when genetically deleted are GPX4 (68, 293, 725), the two genes for cytosolic and mitochondrial TrxR, Txnrd1 (51, 305) and Txnrd2 (123), their cognate substrate Trxs, Txn1 (437) and Txn2 (490), as well as an essential gene for synthesis of GSH (604). Deletion of the Trsp1 gene (encoding selenocysteine (Sec) tRNA, tRNA[Ser]Sec) that is required for synthesis of all selenoproteins is also embryonically lethal (54). Genetic deletion of several other antioxidant enzymes triggers strong phenotypes even if not being embryonically lethal, while certain antioxidant enzymes seem to have an impact only under severe oxidative stress or in specific tissues.

While the SOD family represents the only enzymes able to specifically scavenge O2, catalyzing its disproportionation into O2 and H2O2, multiple classes of enzymes detoxify H2O2 or organic peroxides. Catalases scavenge H2O2 by catalyzing its disproportionation into O2 and H2O. Some selenoproteins and thiol peroxidase such as GPXs, and PRXs, catalyze the two-electron reduction of peroxides to form water using reducing equivalents from GSH or Trx, respectively. The functions of GPXs and PRXs are thus intimately coupled to those of glutathione reductase (GR) and TrxR, enzymes that catalyze the reduction of oxidized GSH (GSSG) and Trx, respectively, using reducing equivalents from NADPH (185, 411). In addition, exciting progress has been made in understanding functions of selenium-dependent methionine-R-sulfoxide reductase 1 (MsrB1) that is a Trx-dependent enzyme. Much progress has also been made in the understanding of selenoprotein P (Sepp1) and Trsp that are both crucial entities for selenium homeostasis and functions of all selenoproteins. Because prior reviews have discussed many of the detailed phenotypes in mice with knockout or overexpression of different antioxidant enzymes (62, 69, 71, 80, 122, 337, 361, 377, 499), we provide herein only an updated synopsis on the same subject, as a basis for the subsequent chapter focusing on their “paradoxical” roles.

A. Superoxide Dismutase Family

1. SOD1

Knockout of Sod1 does not cause embryonic lethality in mice, but results in impairment of the reproduction function of both males and females. Whereas the males produce sperm with decreased motility and fertilizing ability (207, 658), the females display a marked increase in postimplantation embryonic lethality (264) associated with elevated two-cell arrest or cell death (334). The Sod1−/− mice develop anemia (299) and type 1-like diabetes (684). These mice also show a reduced lifespan, a high incidence of hepatocarcinogenesis in late life, and oxidative damage-accelerated spontaneous mutations in liver and kidney (73, 167). Although mice overexpressing SOD1 are apparently normal with a reduced mutation frequency in the cerebellum (357), these animals exhibit certain abnormalities found in patients with Down's syndrome (23, 24, 524, 580).

Knockout and overexpression of SOD1/Sod1 exert negative and positive impacts, respectively, on mouse susceptibility or resistance to neurodegenerative disorders, cerebral and myocardial injuries, diabetic syndrome, and tissue intoxications and dysfunctions (Table 2). However, most, if not all, of the reported “mechanisms” are associations between phenotypes and genetic manipulations, without genuine knowledge of the exact molecular mechanisms that lead to the observed phenotypes. Importantly, the association of SOD1 mutations with familial amyotrophic lateral sclerosis (ALS) does not seem to be due to effects on enzyme activity but rather an increased propensity for protein aggregation (203, 373, 474, 491). Impacts of Sod1 knockout or SOD1 overexpression in mice on neurodegenerative disorders may be related to effects on Aβ oligomerization (477), dopaminergic neurodegeneration (746), lipid peroxidation (536, 647), and protein nitration (294). In cell studies, altering the enzyme affects dopamine autoxidation and changes of GSH status (242), and neuroinflammation driven by activation of nuclear factor-κB (NFκB), release of nitric oxide (NO), and proinflammatory cytokines (153).

Table 2. Physiological impacts or pathological responses of superoxide dismutase-1 overexpression and knockout in mice

Organ/Condition Phenotype Reference Nos.
Overexpression Brain and neurological system Ameliorate brain injuries induced by cold or subarachnoid hemorrhage (in rats) via suppressing MMP-2 and MMP-9 or activation of Akt/GSK-3β 93, 169, 313, 467
Protect vulnerable motor neurons after spinal cord injury via attenuating the mitochondrial apoptosis pathway (in rats) 624, 737
Attenuate kainic acid-induced neurotoxicity in hippocampus and striatum 260, 586
Alleviate phenotypes of Parkinson's disease by elevating dopamine and suppressing lipid peroxidation, protein nitration 294, 536, 647
Vascular system Protect against postangioplasty response and neointimal formation (adenovirus-mediated gene overexpression in rabbit tissue) 358
Lung Alleviate pulmonary oxygen toxicity and prolong survival 695
Resistant to allergen-induced changes in airway control 369
Ischemic injury Protect against cerebral ischemic injury (in mice/rats) 100, 345, 476
Render the heart resistant to myocardial ischemia/reperfusion injuries and protect against ischemia-reperfusion injury, inflammatory responses and apoptosis in cardiac graft 106, 642, 681
Diabetes Protect against diabetogenesis and diabetic nephropathy by suppressing glomerular nitrotyrosine formation and matrix protein synthesis 129, 148, 354
Abolish maternal diabetes-induced embryopathy, block maternal hyperglycemia-induced activation of PKC-α/βII and PKC-δ and lipid peroxidation 232, 391, 679
Cancer Reduce mutation frequency in cerebellum 357
Knockout Brain and neurological system Undergo marked hypertrophy and altered responses to acetylcholine in cerebral arterioles 36, 152
Vulnerable to axonal injury such as axotomy and ischemic insults, and altered calcium homeostasis in spinal motor neurons 556, 611
Increase susceptibility to MPTP-induced phenotypes of Parkinson's disease 746
Drive phenotypes of Alzheimer's disease such as Aβ oligomerization and memory loss 477
Display a modified distribution of fiber types and fiber loss, muscle atrophy and weakness 348, 349, 367
Vascular system Lead to dysfunctions in endothelial-dependent vasodilation and myogenic tone, and accelerated vascular aging in endothelial progenitor cells 125, 151, 152, 228, 674
Lung Increase NFAT activity and NFATc3 nuclear localization resulted from elevated O2/H2O2 ratio, induce spontaneous pulmonary hypertension in pulmonary arteries 546
Liver Alter hepatic gluconeogenesis, glycolysis, and lipogenesis and induce lipid accumulation by impaired lipoprotein secretion 662, 680
Enhance sensitivity to acute paraquat and alcohol-induced liver toxicity 264, 329
Ischemic injury Impair neovascularization induced by hindlimb ischemia 228
Aggravate ischemia/reperfusion-induced myocardial, hippocampal and renal injuries (in mice/rats) 100, 719, 731
Diabetes Accelerate diabetic renal injury 147
Impair islet function, pancreas integrity, and body glucose homeostasis by elevating islet O2, upregulating p53 phosphorylation and downregulating Foxa2/Pdx1 pathway 684
Increase susceptibility to ocular disorders such as cataract and progressive retinal cell loss 247, 295, 500503
Immune response Increase susceptibility to the experimental autoimmune encephalomyelitis 436
Cause anemia and autoantibody production by elevating oxidative stress in erythrocytes 299
Kidney Exhibit an increase in phosphorylation of iron regulatory protein 1(IRP1) in kidney, leading to increased binding to iron-responsive elements (IREs) 734
Susceptible to hydronephrosis- and salt-induced hypertension and histopathological changes 82
Cancer Show reduced lifespan and increased carcinogenesis in late life with oxidative damage-accelerated spontaneous mutations in liver and kidney 73, 167
Others Lead to embryonic two-cell arrest or cell death, and impaired sperm motility and fertilizing ability 207, 334, 658
Induce age-related dysfunction of the lacrimal gland, potentiate hearing loss, cochlear pathology, bone stiffness/strength, skin morphology and wound healing 301, 346, 446, 614

Overexpression of SOD1 protects against various brain and neurological injuries by 1) attenuating the mitochondria-mediated apoptosis pathway (e.g., release of cytochrome c and nuclear translocation of endonuclease G) (624, 737); 2) suppressing the induced expression of matrix metalloproteinases (467); and 3) activating Akt/glycogen synthase kinase 3β (GSK-3β) survival signaling (169, 313). Meanwhile, the protection against cerebral ischemia is conferred in part by upregulating Akt and downregulating p38 mitogen-activated protein kinase (MAPK), and NFκB (100). In contrast, Sod1 knockout potentiates mice to ischemic injuries by activating NFκB (100), and lung dysfunction by increasing nuclear factor of activated T-cells (NFAT) and NFATc3 activities (546). Sod1 knockout can also trigger kidney dysfunction by enhancing the oxidative stress-induced phosphorylation and the conversion of iron responsive protein-1 (IRP1) to the iron responsive element (IRE)-binding form, which may accelerate the reabsorption of iron by renal tubular cells (734). Inhibition of matrix protein synthesis induced by high glucose (129) and the NO-O2 interaction (148) contributes to the protection of SOD1 overexpression against diabetic nephrophathy. Seemingly, several, if not all, of these SOD1-altered phenotypes are associated with specific redox signaling effects, rather than a direct free radical scavenging.

2. SOD2

Sod2−/− mice, unlike Sod1−/− mice, develop cardiomyopathy and neonatal or perinatal lethality, despite variations in postnatal survival time and neuronal injury (370, 392). Thus the reported phenotypes of Sod2 knockout are mostly derived from haplodeficiency or tissue-specific inactivation of the gene. While Sod1−/− female mice become infertile (264), ovaries from postnatal Sod2−/− mice undergo normal folliculogenesis and can produce viable offspring when transplanted to the bursa of wild-type hosts, suggesting the enzyme is dispensable for the ovarian function (443). Interestingly, strain-dependent overexpression of SOD2 is associated with growth retardation and decreased fertility in transgenic mice (542) (Table 3). Although the Sod2+/− mice are viable and no more sensitive to hyperoxia (304), their mitochondria show decreased respiratory capability and elevated induction of the permeability transition (697).

Table 3. Physiological impacts and pathological responses of superoxide dismutase-2 overexpression and knockout in mice

Organ/Condition Phenotype Reference Nos.
Overexpression Brain and neurological system Attenuate MPTP-induced phenotypes of Parkinson's disease 342
Attenuate phenotypes of Alzheimer's disease by reducing hippocampal oxidative stress, modulating Aβ deposition and composition, and slowing memory deficit 164, 435
Lung Prevent hypoxia-mediated decrease in Na+-K+-ATPase and alveolar fluid reabsorption (adenovirus-mediated gene overexpression in rats) 401
Liver Protect against liver mitochondrial DNA depletion and respiratory complex dysfunction after an alcohol binge 433
Diabetes Ameliorate high-fat diet-induced insulin resistance in rat skeletal muscle (electroporation delivery of expression vector to rat muscle) 50
Prevent retinal VEGF expression and retinopathy in diabetic mice 226, 351
Normalize contractility in diabetic cardiomyocytes with improved mitochondrial respiration 597
Ischemic injury Reduce ischemia/reperfusion-induced vascular endothelial cell death and protects against blood-brain barrier damage 425
Reduces neuronal vulnerability to forebrain ischemia (injection of astrocyte-specific expression vector to rat brain) 718
Protect against myocardial ischemia/reperfusion-induced injury 107
Aging Preserve age-associated loss of mitochondrial DNA mass and function of ATP generation 308, 374
Knockout Brain and neurological system Show selective cerebral vascular dysfunction and accelerated disorganization of distal nerve axons following nerve injury 179, 459
Exacerbate phenotypes of Alzheimer disease, Parkinson's disease and ALS 11, 12, 385
Exhibit neurodegenerative phenotypes including frequent, spontaneous motor seizures 188, 394
Liver Exaggerate APAP-induced liver toxicity, mitochondrial dysfunction and DNA fragmentation 200, 545
Diabetes Result in severe central nervous system degeneration and subsequent gait deformities, seizures, and perinatal lethality in type 2 diabetes 497
Heart Induce cardiac mitochondrial dysfunction, severe lipid peroxidation and spontaneous apoptosis in myocardium, and maladaptive cardiac hypertrophy 563, 621, 671
Vascular system Lead to increased vascular oxidative stress with aging and endothelial dysfunction in large and mesenteric arteries 65, 138, 498, 720
Upregulate transferrin receptor and downregulate mitochondrial biogenesis and metabolism in erythroid cells 434
Kidney Develop hypertension, mild renal damage and interstitial inflammation in aged mice 516, 567
Cancer Elevate incidence of neoplasms in aging Sod2+/−Gpx1−/− mice 750
Ischemic injury Increased susceptibility to cerebral ischemia/reperfusion with activation of MMPs, inflammation, blood-brain barrier breakdown and high brain hemorrhage rates 425
Aging Lead to reduced lifespan and premature onset of aging-related phenotypes 653, 675
Others Reduce contractile muscle function and aerobic exercise capacity during aging 335, 417, 418
Ocular pathology including progressive retinal thinning 578
A significant decrease in the respiratory capability and an increased rate of induction of the permeability transition in mitochondria 697

Likewise, knockout and overexpression of Sod2/SOD2 produce negative and positive impacts, respectively, on mouse susceptibility or resistance to a number of acute or chronic disorders (Table 3). Such opposite effects of the enzyme on neurodegenerative disorders are related to regulating mitochondrial ROS generation and function, shifting the amyloidogenic Aβ composition (435), slowing amyloid deposition and memory deficit (164, 435), and modulating dopaminergic neurodegeneration (11, 342). Similarly, the effects on ischemic cerebral injuries are through regulations of blood-brain barrier, matrix metalloproteinases (MMPs), and inflammatory responses (425). Knockout of Sod2 aggravates cellular senescence and aging (653, 675), although overexpression of Sod2/SOD2 fails to extend life span despite preserving age-associated loss of mitochondrial function (308, 374). Overexpression of the enzyme protects against diabetes and complication through improved mitochondrial respiration and integrity and decreased inducible NO synthase and NO production (50, 226, 273, 351, 408, 597). Interestingly, old Sod2+/−Gpx1−/− mice have an elevated incidence of neoplasms (750), suggesting that knockout of multiple antioxidant enzymes can have synergistic effects on carcinogenesis.

3. SOD3

Knockout of Sod3, unlike Sod1 or Sod2, does not affect mouse development and lifespan or further worsen the shortened lifespan of Sod1−/− mice, suggesting limited overlapping roles between these enzymes (81, 593). Respective protective and detrimental outcomes from overexpression and knockout of SOD3/Sod3 are seen in brain, heart and vascular system, kidney, lung, and immune system, as well as in ischemic injuries and carcinogenesis (Table 4).

Table 4. Physiological impacts and pathological responses of superoxide dismutase-3 overexpression and knockout in mice

Organ/Condition Phenotype Reference Nos.
Overexpression Brain and neurological system Protect against brain injury induced by subarachnoid hemorrhage, hyperoxia or cold 447, 510, 739
Improve behavioral outcome from closed head injury 532
Protect against aging-induced memory and cognitive impairments 381, 382
Vascular system Reduce cuff-induced arterial neointimal formation (adenovirus-mediated gene expression in rat tissue) 511
Lung Attenuate pulmonary oxygen toxicity by increasing cGMP activity and reducing of NFκB activation (aerosolized delivery of expression plasmid to neonatal rabbits) or by attenuating neutrophil inflammatory responses 7, 189
Preserve pulmonary angiogenesis by retaining VEGF, VEGFR1, VEGFR2 and PECAM-1 528
Inhibit the development of hypoxia- or fibrosis-induced pulmonary hypertension and vascular remodeling, and ameliorate established pulmonary hypertension 8, 492, 672
Attenuate radiation-, endotoxin (adenovirus-mediated expression)-, influenza-, or air pollutant-induced lung injury 212, 248, 318, 539, 625
Immune response Attenuate inflammatory arthritis by suppressing the production of proinflammatory cytokines and MMPs 736
Ischemic injury Increased resistance to heart or cerebral ischemia/reperfusion injuries 97, 98, 493, 598, 600
Improve recovery from surgical hind-limb ischemia (adenovirus-mediated gene expression) 579
Cancer Inhibit chemical-induced skin carcinogenesis 332
Knockout Heart Exacerbate pressure overload-induced left ventricular hypertrophy and dysfunction 414
Lung Increased susceptibility to hyperoxia 81
Kidney Exhibit renal histopathological abnormalities, hypertension, endothelial dysfunction, and reduced eNOS and Akt activity 326
Immune response Increased susceptibility to the collagen-induced arthritis 570
Ischemic injury Worsen the outcome from cerebral or skeletal muscle ischemia/reperfusion 519, 599

This extracellular SOD isoenzyme plays a pivotal role in protecting against a number of lung disorders including oxidative injury, inflammation, and fibrosis (7, 8, 81, 189, 212, 248, 318, 492, 528, 539, 625, 672). Different from those lacking Sod1, Sod2, or catalase, the Sod3−/− mice are susceptible to hyperoxia and induced oxidative injury (81). Protective roles of the enzyme in pulmonary fibrosis have been thoroughly reviewed (205), and its unique importance in pulmonary function is attributed to its extracellular localization, the pathological importance of extracellular matrix expression, and cytokine release elicited by extracellular ROS. Involved signaling events include modulation of transforming growth factor beta (TGF-β) and early growth response protein 1 (Egr-1) expression (672), preserving angiogenesis (528), and maintaining NO bioavailability and subsequently modulating cGMP and NFκB activity (7). The unique protection by SOD3 against lung oxidative insults offers the potential of administrating the enzyme to relieve pulmonary disorders (205). Moreover, the distribution of SOD3, rather than the total SOD activity, in the extracellular space is crucial for protecting the heart against the pressure overload as this insult renders the Sod3−/− mice with elevated myocardial O2 production and nitrotyrosine formation, increases of ventricular collagen I and III as well as MMP-2 and -9, and decreases in ratio of GSH/GSSG (glutathione disulfide) (414). Knockout of Sod3 renders mice susceptible to collagen-induced arthritis (570), while overexpression of SOD3 in mouse synovial tissue attenuates inflammatory arthritis (736), via opposite modulations of the production of the proinflammatory cytokines such as interleukin (IL)-1β and tumor necrosis factor (TNF)-α as well as MMPs. Knockout of Sod3 also impairs renal-vascular function, in part by decreasing Akt and eNOS phosphorylation and heme oxygenase 1 activity (326).

B. Catalase

Catalase is ubiquitously expressed and is predominantly located in peroxisomes of all types of mammalian cells with the exception of erythrocytes (669) and human vascular cells (607). A certain activity of catalase is also detected in mitochondria of rat heart (541). In humans, acatalesemia is a comparatively common genetic disease with near-total lack of catalase, which is typically considered to be asymptomatic but may be associated with increased risk of a number of diseases (224). Cat−/− mice show normal development and fertility (268) and are not more susceptible to hyperoxia-induced lung injury than the wild-type controls. Overexpression of catalase in mitochondria prolongs the lifespan of mice and attenuates age-associated pathological changes (137, 504, 584, 654). The overall outcomes of Cat knockout are rather limited, especially compared with those associated with Sod. This may in part be due to the fact that GPX and PRX (90, 185) play major roles in removing H2O2 at relatively low concentrations in the cells, whereas the contribution of catalase increases when intracellular H2O2 is high (428).

In contrast, promotion of health by CAT/Cat overexpression has been shown in many tissues and conditions (Table 5). Overexpression of CAT protects against cardiovascular injuries or dysfunction (424, 723, 724), which is particularly relevant due to the lack of the enzyme activity in the human vascular smooth muscle and endothelial cells (607). Aortas from apolipoprotein E knockout mice overexpressing CAT show smaller and relatively early stages of atherosclerotic lesions compared with the control (722). Cardiac-specific overexpression of rat Cat attenuates the paraquat-induced myocardial geometric and contractile alteration by alleviating JNK-mediated endoplasmic reticulum stress (208), prolongs lifespan, and suppresses aging-induced cardiomyocyte contractile dysfunction and protein damage (712). Also, this specific overexpression of rat Cat rescues the anthrax lethal toxin- or lipopolysaccharide (LPS)-induced cardiac contractile dysfunction by alleviating oxidative stress, autophagy, and mitochondrial injury (317, 660). The elevated catalase antagonizes the alcohol dehydrogenase-associated contractile depression after acute ethanol exposure in murine myocytes, partially through improving intracellular Ca2+ handling and ablation of alcohol dehydrogenase-amplified JNK activation and Erk deactivation (748, 749). Comparatively, the endothelium specific overexpression of CAT shows a weak protection against myocardial or vascular ischemia/reperfusion injury, despite preserving the responsiveness of the heart to adrenergic stimulation (704). Knockout of Cat accelerates diabetic renal injury through upregulation of TGF-β and collagen secretion (289), whereas overexpression of Cat protects against the pathogenesis via attenuation of angiotensinogen and Bax function and normalized expression of angiotensin converting enzyme 2 (ACE-2) (58, 603).

Table 5. Physiological impacts and pathological responses of catalase overexpression and knockout in mice

Organ/Condition Phenotype Reference Nos.
Overexpression Heart Preserve the responsiveness of the heart to adrenergic stimulation 704
Attenuate cardiac contractile dysfunction induced by paraquat, anthrax, LPS, acute ethanol, or hypoxia-reoxygenation, through alleviating events such as JNK-mediated ER stress 105, 208, 317, 660, 748, 749
Prevent progressive myocardial remodeling, including myocyte hypertrophy, apoptosis, and interstitial fibrosis due to overexpression of Gαq 538
Protect against acute and chronic doxorubicin-induced cardiotoxicity 319, 320
Vascular system Prevent pathological mechanical changes underlying abdominal aortic aneurysm formation 424
Reduce pressure response to norepinephrine or angiotensin II by eliminating H2O2 in arterial wall 723
Inhibit toxin-accelerated atherosclerosis in the hypercholesterolemic ApoE−/− mice 722, 724
Kidney Inhibit the development of hypertension and renal injury in the angiotensinogen transgenic mice 219
Ischemic injury Render the heart resistant to myocardial ischemia/reperfusion injury 386
Diabetes Increase resistance to STZ-induced β-cell injury and diabetic effect including an attenuation of renal angiotensinogen and proapoptotic gene expression 58, 99, 717
Cardiac overexpression rescues insulin resistance-induced cardiac contractile dysfunction 160
Prevent hypertension and progression of nephropathy by attenuating renal oxidative stress and normalizing ACE-2 expression in type 1 diabetes 603
Aging Mitochondrial overexpression prolongs lifespan with delayed age-associated pathologies including cardiac aging and cataract, decreases malignant nonhematopoietic tumor burden in old mice and enhances hippocampus-dependent memory with a reduction of anxiety 137, 504, 584, 654
Cardiac overexpression prolongs lifespan and attenuates aging-induced cardiomyocyte contractile dysfunction and protein carbonyl formation 712
Knockout Brain Show a decreased efficiency in brain mitochondrial respiration following cortical oxidative injury 268
Kidney Render remnant kidneys increased susceptibility to oxidant tissue injury and progressive renal fibrosis after nephrectomy 344
Diabetes Accelerate STZ-induced diabetic renal injury with increased expression of glomerular TGF-β and collagen α1 289

C. GPX Family

GPX enzymes utilize reducing equivalents from GSH to reduce peroxides (60, 94, 185). Eight isoforms of GPX are known, of which five are selenoproteins (GPX1-4 and GPX6). The three selenium-independent GPX enzymes rely on thiol rather than selenol chemistry. Among the GPX enzymes, GPX1 is the most abundant and ubiquitous isoform. GPX6 is found as a selenoprotein only in humans, while the orthologous Gpx6 has a catalytic cysteine (Cys) in mice and several other species (353). Several recent reviews have summarized physiological roles of GPX enzymes in relation to other selenoproteins (6062, 124, 185, 360, 361, 554). Our discussion herein mainly summarizes the key findings from genetic mouse models.

1. GPX1

Gpx1 is not essential for survival or reproduction, despite its protection against cataract and slight growth retardation (110, 141, 172, 265). Knockout of Gpx1 sensitizes mice to pro-oxidant-induced oxidative injuries, whereas overexpression of the enzyme confers extra protection against such injuries in various tissues (Table 6). The elevated susceptibility of the Gpx1−/− mice to various acute oxidative injuries, including increased lethality induced by high doses of paraquat and diquat, relates to accelerated oxidation of NAD(P)H, proteins, and lipids (108, 111, 141, 195, 196). The importance and mechanism for Gpx1 protection depends on the intensity of stress as well as antioxidant status of the challenged animals (108, 111, 141, 195, 196, 376), but high levels of dietary vitamin E do not replace protection of Gpx1 (113). It should also be noted that Gpx1 is one of the most selenium-responsive selenoproteins, whereby low dietary selenium intake rapidly lowers its expression in most tissues (631), suggesting that the selenium status of the control mice will affect the comparative outcome of Gpx1 removal.

Table 6. Physiological impacts and pathological responses of glutathione peroxidase-1 overexpression and knockout in mice

Organ/Condition Phenotype Reference Nos.
Overexpression Brain and neurological system Protect against 6-OHDA-induced neurotoxicity, trauma-induced mitochondrial dysfunction, cerebral/ischemia reperfusion and hypoxic ischemic injury 42, 564, 595, 691, 716
Heart Resistant to ischemia/reperfusion injury and doxorubicin-induced cardiomyopathy and mitochondrial dysfunction 714, 733
Liver Protect against paraquat-induced hepatotoxicity 111
Enhanced susceptibility to acetaminophen toxicity 458
Diabetes and metabolic disorders Contribute to hyperinsulinemia in association with the transcription factor Pdx1 and mitochondrial uncoupling protein 2 686
β-Cell-specific overexpression of human GPX1 rescues β-cell dysfunction and reverses diabetes in 20-wk-old db/db obesity mice 229, 244
Gpx1 overexpression mice are obese 445
Knockout Brain and neurological system Exacerbate neuronal toxicity induced by Aβ, malonate, 3-nitropropionic acid, and 1-methyl-4-phenyl-1,2,5,6-tetrahydropyridine 128, 341
Resistant to kainic acid-induced seizure and neurodegeneration 309
Heart Susceptible to ischemia/reperfusion injury in male mice 397, 732
Mutate the benign coxsackievirus B3 and induce myocarditis 37
Susceptible to doxorubicin- and angiotensin II-induced aortic and cardiac dysfunction and oxidative stress 15, 206
Accelerated progression of atherosclerosis under a diabetic ApoE−/− background on a high fat diet (21% fat, 0.15% cholesterol) 384, 652
Liver and kidney Enhanced susceptibility to diquat- and paraquat-induced oxidative stress 110, 112, 114, 195, 199
Resistance to peroxynitrite-mediated hepatic toxicity 199
Accelerated diabetic nephropathy in the ApoE−/− model of diabetes 115, 640, 641
Lung Enhanced oxidation and lung inflammation after cigarette smoking or influenza A infection 165, 726
Aging No apparent phenotype except for cataract 142, 265
Carcinogenesis Gpx1−/−Gpx2−/− mice have spontaneous polyps formation and inflammation-induced tumor formation in the gastrointestinal tract 174

Protections against disease conditions by GPX1 are illustrated by increased susceptibility of Gpx1−/− mice and resistance of Gpx1 overexpressing mice to various oxidative insults, including ischemia/reperfusion and hypoxic ischemic injury in the brain, heart, and liver (128, 195, 341, 397, 564, 595, 691, 732). Furthermore, Gpx1 protects against cardiomyopathy induced by coxsackievirus B3 through suppression of viral genome mutation (37), atherosclerosis in a pro-diabetic ApoE−/− mouse model (115, 384, 652), doxorubicin-induced and angiotensin II-mediated functional declines and cardiac hypertrophy (15, 206, 714, 733), defective blood flow and epithelial progenitor circulation in a model of ischemia-induced angiogenesis (202), diabetic nephropathy in association with fibrosis and inflammation (115, 640, 641), and detrimental effects of cigarette smoking or influenza A infection in the lung (165, 726).

Although the exact molecular mechanisms for involvement of Gpx1 in the above-described pathogeneses are largely unknown, the existing evidences point to ROS scavenging and redox signaling as the main modes of action. The Gpx1−/− mouse brain shows elevated oxidative stress, caspase-3 cleavage, and 3-nitrotyrosine formation (128, 341). The decreased migration of endothelial progenitor cells in the Gpx1−/− mice toward vascular endothelial growth factor (VEGF) and capability of these cells in promoting the formation of vascular network are indeed related to the elevated intracellular ROS levels (202). The protection of Gpx1 against diabetic nephropathy is associated with decreases of hydroperoxides, 8-isoprostane, nitrotyrosine, 4-hydroxynonenal, and proteins implicated in fibrosis and inflammation (115, 640, 641).

2. GPX2

GPX2 was first found in the gastrointestinal tissues (117). There are no Gpx2 transgenic mouse lines reported (Table 1). Like the Gpx1−/− mice, Gpx2−/− mice appear normal unless they are stressed by oxidative challenges (678) (Table 7). The Gpx1 expression is upregulated in the colon and ileum of Gpx2−/− mice (186), which may explain why they do not develop cancer spontaneously but develop squamous cell tumor when additional stress such as ultraviolet exposure is employed (678). Likewise, spontaneous polyps are developed in Gpx1−/−Gpx2−/− mice, probably due to elevated intestinal lipid peroxidation with onset of inflammatory bowel disease (118, 174) (Table 6). Notably, nuclear factor (erythroid-derived 2) (NF-E2)-related factor (NRF2), a redox-sensing transcription factor, may counteract oxidative injuries partially through upregulation of Gpx2, at least in lung (613). Given its high expression in the gastrointestinal tract, GPX2 likely exerts antioxidant or antitumorigenic functions there, in association with GPX1 and NRF2. However, a basic question still remains as to whether knockout of Gpx2 itself elevates intracellular H2O2 levels or affects NRF2 (352).

Table 7. Physiological impacts and pathological responses of glutathione peroxidases 2–4 overexpression and knockout in mice

Organ/Condition Phenotype Reference Nos.
Overexpression Heart Mitochondrial Gpx4 (rat) attenuates ischemia/reperfusion cardiac injury 136
Liver Human GPX4 protects against diquat-induced apoptosis and oxidative stress 393, 548
Increased expression of human GPX3 in the plasma by 50% renders the mice resistant to APAP-induced hepatotoxicity and a thermosensitive phenotype 457, 458
Knockout Brain and neurological system Gpx3/−: display cerebral infarctions 311
Gpx4+/−: accumulate oxidized lipids and senile plagues 101
Mitochondrial Gpx4−/−: apoptosis-induced cerebral degeneration in the hindbrain 52
Gpx4−/− (neuron-specific): neurodegeneration, corrected by α-tocopherol 590
Gpx4−/− (endothelium-specific): vitamin E-dependent suppression of angiogenesis in aortic explants 707
Nuclear Gpx4−/−: retardation in atrium formation 52
Carcinogenesis Gpx2−/−: severe inflammation and colon carcinoma induced by AOM/DSS; prone to UV-induced squamous cell tumor 352, 678
Gpx3−/−: prone to colitis-associated carcinoma and increased inflammation in the colon 32
Aging Gpx4+/−: a slight lifespan extension (1,029 vs. 963 days) 549
Others Gpx4−/− (photoreceptor-specific): degeneration and apoptotic death of photoreceptor cells 663
Gpx4−/− (spermatocyte-specific): infertility; reduced forward mobility and mitochondrial membrane potential in the spermatozoa 292
Mitochondrial Gpx4−/−: infertility; impaired sperm quality and severe structural abnormalities in the midpiece of spermatozoa 581

3. GPX3

GPX3 is mainly synthesized in proximal convoluted tubule cells of the kidney (22). While the majority of renal GPX3 is secreted into plasma, some retains at the basement membranes to account for 20% of total selenium in kidneys (429, 505). Independent of its peroxidase activity, this enzyme transfers selenium from the dams to the fetus (72), while Sepp1 instead of Gpx3 provides selenium to neonates via the milk (257). Knockout of Gpx3 and overexpression of GPX3 in mice produce essentially opposite impacts on ROS-related events (Table 7). The Gpx3−/− mice display cerebral infarctions, along with elevated oxidative stress, blood clots, the induction of P-selectin, and lowered plasma cGMP levels (311) and colitis-associated carcinoma with increased inflammation in the colon (32). Overexpression of GPX3 renders mice resistant to acetaminophen (APAP) overdose (458) but leads to hyperthermia (457). Thus some physiological effects of Gpx3/GPX3 modulation can be viewed as unexpected if the enzyme would solely have a role in extracellular H2O2 scavenging. It is therefore possible that it has yet unrecognized physiological functions that are not directly related to the extracellular enzymatic activity.

4. GPX4

GPX4 has three isoforms in the cytosol, mitochondria, and sperm nucleus and enzymatically exhibits substrate preference toward phospholipid hydroperoxide (667). Interacting with guanine-rich sequencing-binding factor 1, GPX4 suppresses lipid peroxidation and apoptosis during embryogenesis (664). Because the global knockout of Gpx4 renders embryonic lethality, tissue-specific and Gpx4 isoform-specific conditional knockout mice have been generated (Table 7). Collectively, increased levels of lipid peroxides by localized Gpx4 deficiency lead to 1) endothelial cell death and thrombus formation in a vitamin E-dependent manner (707); 2) 12/15-lipoxygenase-dependent apoptosis-inducing factor (AIF) translocation and neuronal apoptosis (590); 3) decline of mitochondrial potential and infertility of spermatozoa (292); and 4) defective photoreceptor maturation (663). Recently it was shown that cell death by ferroptosis is triggered upon genetic removal of Gpx4 in either kidney (194) or T cells (441). Clearly, Gpx4 is important for protections against the detrimental effects of lipid peroxidation, but the enzyme also has an intriguing peroxidase-independent structural role in sperm maturation (667).

Results from isoform-specific knockout of Gpx4 indicate that 1) mitochondrial Gpx4 protects against apoptosis during hindbrain development (52); 2) mitochondrial Gpx4 suppresses protein thiol content and is essential for male fertility (581); and 3) nuclear Gpx4 is essential for atrium formation (52), but indispensable for sperm maturation (581). Because the mitochondrial or nuclear Gpx4−/− mice are viable, the cytosolic Gpx4 confers the embryonic lethality phenotype of the global Gpx4 knockout. Reciprocally, overexpression of GPX4/Gpx4 in the global Gpx4−/− mice, detected only in liver and heart, can rescue their embryonic lethality and attenuate the induced decline of mitochondrial potential (393, 548) (Table 7). Similarly, the mitochondrion-specific Gpx4 overexpression maintains mitochondrial membrane potential and protects against ischemia/reperfusion in the heart (136).

D. TrxR Family

TrxRs are a family of NADPH-dependent selenoproteins that play important roles as key propagators of the Trx system and thus several Trx-dependent enzymes, including PRXs, Msrs, ribonucleotide reductase (RNR), sulfiredoxin, and more (17, 122, 412, 423, 571). Three mammalian genes encode different TrxR isoforms, in mice being Txnrd1 encoding cytosolic TrxR1 (215, 325, 507), Txnrd2 encoding mitochondrial TrxR2 (also called TR3) (325, 454, 565, 628), and Txnrd3 encoding thioredoxin glutathione reductase (TGR) that is mainly expressed in spermatids of the testis and seems to be important for spermatogenesis (211, 623, 626, 627, 659).

All Txnrd genes are transcribed in a complex manner, resulting in divergent forms of each isoenzyme that differ from each other mainly in their NH2-terminal domains (88, 95, 139, 211, 442, 455, 507, 572, 573, 622, 629), potentially reflecting many levels of regulation. The phenotypes of mouse knockout models targeting the Txnrd1 and Txnrd2 genes are summarized in Table 8. No knockout models targeting Txnrd3 have yet been reported, and overexpression of TrxR isoenzymes is difficult to obtain, due to their intricate expression patterns.

Table 8. Physiological impacts and pathological responses of thioredoxin reductases knockout in mice

Organ Condition Phenotype Reference Nos.
Txnrd1 (TrxR1, TR1)
Knockout Ubiquitous deletion Early embryonic lethality 51, 305
Heart (MLC2a-driven knockout) No apparent phenotype and no effect on infarct size after cardiac ischemic/reperfusion injury 280, 305
Nervous system (Nestin-driven knockout) Smaller mice with ataxia and tremor, cerebellar hypoplasia, ectopically located and abnormal Purkinje cells, disorganized Bergmann glial network 617
Neurons (Tα1-driven knockout) No apparent phenotype 617
Liver (Alb-driven knockout) Strong upregulation of Nrf2-targeted genes, no apparent growth defect in nontreated liver but metabolic switch with accumulation of glycogen or lipids and significantly increased resistance to acetaminophen challenge 302, 520, 535, 634
B-cell lymphoma (mb-1-driven knockout in λ-myc lymphoma model) No effect on tumor growth except induction of an absolute requirement of the tumors on GSH upon Txnrd1 knockout 430
Hepatocellular carcinoma (induced by diethylnitrosamine in Alb-driven knockout) Strongly increased propensity for hepatocarcinogenesis 79
Txnrd2 (TrxR2, TR2)
Knockout Ubiquitous Early embryonic lethality with impaired heart development, anemia, lack of hematopoiesis and liver apoptosis 123
Nervous system (Nestin-driven knockout) No apparent phenotype 617
Heart (MLC2a-driven knockout) Congestive heart failure with signs of dilated cardiomyopathy, death within hours after birth 123
Heart (tamoxifen-induced α-myosin heavy chain-driven knockout) Impaired cardiac function at rest and more severe injuries after cardiac ischemia/reperfusion, with NAC treatment normalizing the observed phenotypes 280
B- and T-cells (CD4- and CD19-driven knockouts) No apparent phenotype 209

1. TrxR1

The full Txnrd1−/− knockout mice display early embryonic lethality, with one study reporting lethality between embryonic days 8.5 and 10.5 mainly due to decreased cellular proliferation (305), and the other study embryonic death before day 8.5 with a lack of formation of mesoderm (51). Differences in genetic targeting between these studies, one removing the last exon of the gene (305) and the other removing the first exon (51), may possibly help to explain the different phenotypes. Notably, the knockout in mice of the Trx1/Txn gene encoding Trx1 (see below) that is the presumed main substrate of TrxR1, gives even earlier embryonic death than upon TrxR1 removal (437). This suggests that functions of TrxR1 and Trx1 are not always directly linked in a physiological setting, which may be due to the fact that the GSH system can also keep Trx1 reduced through Grx activities (162). It is, however, clear that TrxR1 is an essential enzyme for embryonic development in mice.

Heart-specific Txnrd1−/− mice are normal (305), as are mice with neuron-specific deletion of the enzyme (617). Interestingly, however, expression of the enzyme in glial cells is essential for normal development of the central nervous system (617). When deleted in either hepatocytes, mouse embryonic fibroblasts or B-cell lymphoma cells, the Nrf2-driven and mainly GSH-dependent enzyme systems are typically strongly upregulated (302, 430, 520, 535, 634). In fact, it was found that the Nrf2 induction can be so strong upon TrxR1 deletion or inhibition that cells become even more resistant to certain events of oxidative challenge than those having normal expression of TrxR1 (63, 302, 405, 634). These apparently paradoxical impacts on mouse susceptibility to stress upon TrxR1 removal will be further elaborated in the following sections.

2. TrxR2

Similarly to TrxR1, the mainly mitochondrial isoenzyme TrxR2 is essential for embryonic development. Interestingly, however, Txnrd2 knockout yields early embryonic death in a more tissue specific manner, presenting liver apoptosis, impaired hematopoiesis, and insufficient heart development (123). Knockout of mitochondrial Trx2 that is presumed to be the main substrate of TrxR2, however, displays a more severe phenotype with massive widespread apoptosis and open anterior neural tube (490). This illustrates that the functions of TrxR2 are not always directly linked to those of Trx2, which is similar to the situation with Trx1 and TrxR1 (see above).

There was a lack of overt phenotype when TrxR2 was conditionally knocked out in the nervous system (617) or in B- and T-cells (209), while its conditional knockout in heart produced obvious detrimental effects (123, 280). Recently, it was also shown that TrxR2 knockout in tumor cells prevented tumor growth because of a lack of hypoxia-inducing factor (HIF) function and JNK activation (254). These observations suggest that although most cells and tissues are dependent on mitochondrial function, the physiological effects of genetic deletion of the mitochondrial TrxR2 enzyme are more specific than what would be explained by a generally impaired mitochondrial function in the whole organism.

E. Additional Mouse Models for Knockouts of Selenoproteins

Most, if not all, of the 24-25 selenoproteins in the mammalian proteomes (353) presumably have redox activity. Readers are referred to other recent reviews for a full survey of these proteins (84, 124, 250, 322, 360). However, in the context of this article it is worth considering MsrB1, a Trx1-dependent selenoenzyme, and Sepp1, as their physiological antioxidant roles have been studied using several genetic mouse models. Knockout of MsrB1 renders mice prone to lipid peroxidation and protein oxidation in tissues as well as defective actin polymerization in macrophages upon LPS challenge (190, 371) (Table 9). Neuronal protection by Sepp1, a predominant extracellular selenoprotein that delivers selenium from liver to other tissues and has peroxidase activity (576, 639), may be attributed to its selenium transport function, because deletion of its COOH-terminal region being rich in Sec residues (amino acids 240–361) was sufficient to produce severe neurodegeneration in mice (258, 533, 583). Liver-specific expression of SEPP1 in Sepp1−/− mice enhances their brain selenium content and rescues the neurological defects (559), further supporting the important role of this selenoprotein in the selenium transport.

Table 9. Physiological impacts and pathological responses of overexpression, knockout, and transgene of other selenium-dependent proteins and Trsp in mice

Models Phenotypes Reference Nos.
Trsp transgene Lacking the STAF-binding site Tissue-specific decrease in selenoprotein expression (brain, muscle > lung, spleen > liver, kidney; no change in heart and testes) 77
Mutation at position 37 (A→G) Tissue- and selenoprotein-specific changes in selenoprotein expression (↓Gpx1, ↑TrxR1, liver > testes; pyogranulomatous inflammation in various tissues] 470, 471
mTOR-dependent increase of muscle growth after exercise 279
Severe neurological defects and mortality in these mice on a selenium-deficient or high selenium (2.25 ppm) diet 324
Increased susceptibility to azoxymethane-, diethylnitrosamine-, and C3(1)-induced carcinogenesis in intestines, liver, and prostate, respectively; no changes on TGF-α-induced hepatocarcinogenesis 159, 296, 324, 470
Increased X-ray-induced micronuclei formation in the erythrocytes 27
Defective immune responses after the lungs were targeted with viral infection of influenza 601
Trsp conditional knockout Hematopoietic cells (Mx1-driven) Prone to hemolytic anemia and defective oxidative homeostasis 327
Macrophage (LysM-driven) Increased oxidative stress, induction of Nrf2 expression, defective immune response and expression of fibrosis-associated genes 635
T cells (Lck-driven) Defective T-cell maturation and antibody responses upon T cell receptor stimulation 608
Neuron (Tα1-driven) Defects in interneuron development and cerebellar hypoplasia, increased striatal neuronal loss with movement disorder, and seizure due to spontaneous epileptiform activity 588, 701, 702
Endothelial cells (TieTeK2-driven) Embryonic lethality, 14.5 days embryos being smaller with underdeveloped vascular system, limbs, tail and head 609
Osteo-chondroprogenitor (Col2a1-driven) Growth retardation and delayed skeletal ossification reminiscent of Kashin-Beck disease 161
Skin (K14-driven) Small body size, alopecia, flaky and fragile skin, and early regression of hair follicles 592
Heart and skeletal muscle (MCK-driven) Die 12 days after birth with acute myocardial failure 609
Mammary gland (MMTV- or Wap-driven) Increased mammary carcinogenesis 288
Liver (Alb-driven) Premature death at 1–3 mo of age, no changes in brain selenium levels, and increased apolipoprotein E and cholesterol levels in plasma 76, 587, 591
Kidney (NPHS2-driven) No effect on streptozotocin-induced diabetes 48
Prostate epithelium (ARR2PB-driven) Early onset of intraepithelial neoplasia 415
Trsp variant transgene Mutant G37 transgene under global Trsp−/− Reduced fertility in males and litter size in females 78
A34 or G37 transgene in liver-specific Trsp−/− Reversal of the elevated levels of apolipoprotein E and cholesterol in the plasma 591
Other selenoprotein knockout Sepp1−/− Neuronal degeneration, loss of 55% Se in the brain, degenerated and dystrophic axons in the cervical spinal cords and the brainstem 259, 533, 583
Sepp1Δ240–361 Decreased Se levels in the brain 258
MsrB1−/− Oxidation of protein, lipid, and GSH in liver and kidney; actin fragmentation 190, 371

The redox activity of all selenoenzymes depends on the function of Sec, which is cotranslationally incorporated at redefined specific UGA codons in a process that requires tRNA[Ser]Sec, the transcriptional product of the Trsp gene. Because Trsp−/− mice are embryonically lethal (54), various conditional knockouts and variants of Trsp have been made to study roles and regulations of selenoproteins in specific tissues, resulting in several interesting phenotypes (Table 9). Intriguingly, knockout of Trsp in endothelial cells causes embryonic lethality and in muscle and liver induces postnatal death (76, 609). Global or conditional Trsp−/− mice expressing wild-type or mutant Trsp transgene have also been generated (78, 591). These Trsp-altered mouse models help understand tissue-specific functions of selenium and allow for recapitulation of mechanisms behind the classical selenium-deficiency syndrome Kashin-Beck disease (161). A recent review (361) offers detailed discussion on the pleiotropic effects of Trsp targeting that are likely to be derived from the combined effects of modulation of multiple selenoproteins at once.

F. Thioredoxins and Glutaredoxins

Trxs are small thiol-disulfide oxidoreductases with a Cys-Gly-Pro-Cys active site and are present in all living cells. The reduced forms with a dithiol motif in the active site catalyze disulfide reduction reactions, generating oxidized forms of Trx with a disulfide in the active site, which is again reduced by NADPH via TrxRs (276, 396). The isoforms of Trx have a broad range of functions in mammalian cells (18), including serving as electron donors for Prxs that are controllers of the intracellular redox state together with GSH (274), and being major protein S-denitrosylases (91). The structure of Trxs comprises a central core of β-strands surrounded by α-helices that defines the Trx-fold, now known to be present in a large number of proteins denoted the Trx superfamily of proteins. This includes Grx (395), glutathione S-transferases, GPXs, PRXs, and proteins of the protein disulfide isomerases (PDI) family, which are all built from Trx domains (21). In the context of this review, the main results of genetic mouse experiments for analyses of Trx1, Trx2, Grx1, and Grx2 are discussed as follows.

1. Trx1

Trx1 (encoded by Txn in mice) is ubiquitously expressed in the cytosol/nucleus and has a large number of functions in cellular redox control and antioxidant defense (18). One of those functions is to provide reducing power to RNR that is essential for DNA synthesis. Because the global knockout of Txn in mice induces early embryonic lethality (437), shortly after implantation with differentiation and morphogenesis defects, studies in adult mice were instead enabled using a dominant-negative mutant line in which the active site Cys-32 and Cys-35 residues were altered to Ser (dnTrx-Tg) (140). These functionally Trx1-deficient mice display decreased Trx activity in the lung and are sensitive to ambient air at room temperature. These mice experience genotoxic stress, as evidenced by decreased activities of aconitase and NADH dehydrogenase, lower mitochondrial energy production, but increased levels of p53 and Gadd45α expression. These dnTrx-Tg mice are also manifested with increased levels of proinflammatory cytokines (140), which are aggravated by exposure to hyperoxia. In contrast, overexpression of enzymatically active Trx1 in the lung (140) helps maintain redox balance and mitochondrial function with decreased inflammation. Mice overexpressing TXN have increased resistance to a range of oxidative stress insults (643). In addition, Trx1 has been shown to protect against joint destruction in a murine model of arthritis (657). Overexpression of the protein furthermore seems to promote fetal growth by reducing oxidative stress in the placenta (665), prevent diabetic embryopathy (314), and extend mainly the earlier part of the life span in mice with a prolonged youth phenotype (527).

Trx1 is secreted from cells under inflammation and oxidative stress and is detectable in plasma (482). Of particular interest is that the extracellular Trx1 is taken up by cells and has been proposed as an effective antioxidant therapy (439, 483, 688). Its presumed antioxidant and antiapoptotic properties are tightly coupled with the reduced form of Trx1 binding to thioredoxin interacting protein 1 (TXNIP) (475, 735) or apoptosis signal-regulating kinase (ASK1, also called MAP3K5) (290, 310). Extracellular Trx1 is however also found in plasma as a truncated form called Trx80, resulting from α-secretase cleavage (213) and known to act as an inflammatory mediator (Th1) via effects on the immune system and monocytes (522). Both Trx1 and Trx80 seem to have a positive effect protecting from Alzheimer's disease in the brain (213). Because Trx80 lacks redox activity together with TrxR1 (523) and since extracellular forms of these proteins are likely to remain oxidized, it is possible and even likely that some of their physiological roles are unrelated to redox activities.

2. Trx2

Trx2, with a mitochondrial leader sequence, is targeted to the mitochondria, where it plays a crucial role in controlling ROS by acting as a reductant of Prx3 in concert with the GSH system and Grx2 (240, 744). Knockout of the Trx2/Txn2 gene (490) induces embryonic lethality with massively increased apoptosis and exencephaly with open anterior neural tube. Cardiac specific deletion of Trx2 (283) produces spontaneous dilated cardiomyopathy at 1 mo of age, with increased heart size, reduced ventricular wall thickness, and progressive decline in left ventricular contractile function resulting in mortality due to heart failure at young age. In cardiomyocyte-specific Tnx2−/− mice, mitochondrial function and ATP production are declined and ASK1-dependent apoptosis accelerated. Interestingly, humans with dilated cardiomyopathy have lowered Trx2 protein levels in heart tissue, suggesting that these mice could be a good model of the human disease (283).

3. Grx1

Grx1 catalyzes GSH-disulfide oxidoreduction reactions (275), deglutathionylation of S-gluthionylated proteins (277), and reduction of Trx1 by GSH when TrxR is inactivated (162). Surprisingly, knockout of Grx1 (also named Glrx) (267) results in only a mild phenotype without major effects on ischemia reperfusion injuries. However, knockout of the gene offers protection against inflammation or defective revascularization in diabetes (4, 270), which will be further elaborated on in the following section.

4. Grx2

Grx2 is encoded by a gene resulting in splice variants including Grx2a located in mitochondria and Grx2c in the cytosol/nucleus. Knockout of Grx2 (also named Glrx2) (710) induces early onset of age-dependent cataract in mice. Grx2 is also required to control mitochondrial function since knockout affects cardiac muscle (426, 427), giving rise to larger hearts and high blood pressure.

G. Peroxiredoxin Family

The PRX enzymes are a family of abundantly present 20–30 kDa peroxidases (185, 562, 706). These homodimeric proteins fall into three varieties distinguished by their reaction mechanisms and the number of cysteine residues required for catalysis: typical 2-Cys (in mammals PRX1-4), atypical 2-Cys (mammalian PRX5), and 1-Cys (mammalian PRX6) (561, 594, 706). Both types of 2-Cys PRX utilize the reducing power of NADPH via the Trx/TrxR system to reduce their active site disulfides, formed upon catalysis with peroxide reduction, back to active dithiols. On the other hand, 1-Cys PRXs mainly utilize GSH as the reducing agent (706). Furthermore, the various PRX isoforms exhibit different subcellular localizations (271, 706). As the PRX enzymes are highly abundant, accounting for as much as 1% of soluble cellular protein (673, 706) and are excessively reactive with H2O2, they are likely to be critical for both oxidative stress protection as well as redox signaling (562, 616, 698, 699).

1. Effects of Prx knockout

Mice lacking Prx1-4 and 6 are viable, but exhibit increased ROS levels and sensitivity to oxidative insults (300, 375, 389, 461, 484, 683). In general, both Prx1−/− and Prx2−/− mice appear healthy and are fertile, but have hemolytic anemia and increased atherosclerotic plaques (338, 518), suggesting that Prx1 and Prx2 protect red blood cells from oxidative stress (375, 484). Indeed, they exhibit splenomegaly, Heinz bodies in their blood, and morphologically abnormal red blood cells, which are high in ROS (375, 484). Prx3−/− mice are healthy in appearance and could grow to maturity, but exhibit elevated intracellular ROS, including in lung tissue (389). Intratracheal inoculation of LPS to the Prx3−/− mice results in pronounced lung inflammation (389). Likewise, Prx6−/− mice also appear normal, but are very sensitive to oxidative insults (461, 683).

2. Effects of Prx overexpression

Overexpression of the PRX enzyme genes generally confers protection against different forms of oxidative stress. For instance, overexpression of Prx3 in heart mitochondria of mice suppressed cardiac failure after myocardial infarction (440). In addition, the Prx3 overexpressing mice have lower mitochondrial H2O2 concentrations and are protected against hyperglycemia and glucose intolerance (102). Overexpression of Prx2 inhibits the ischemic damage of neurons (56). Overexpression of PRX4 in the pancreas of mice suppresses the TRAIL-mediated apoptosis, protects pancreatic islet β-cells against injury caused by single high-dose streptozotocin (STZ)-induced insulitis, and attenuates inflammation (155). When Prx6 is overexpressed, development of cataracts in mouse and rat lenses are significantly delayed (355). These transgenic mice exhibit extra resistance to the lung injury induced by hyperoxia (687). However, global Prx6 overexpression does not protect against diet-induced atherosclerosis despite lowering levels of H2O2 (531).

III. “PARADOXICAL” OUTCOMES

Although knockout of several antioxidant enzymes is detrimental and their overproduction beneficial to health, the opposite impacts have also been increasingly observed. This section describes a series of such apparently “paradoxical” cases that reveal metabolic benefits of deleting major antioxidant enzymes, or harmful effects of overexpressing them.

A. SOD Family

1. Elevated resistance to APAP toxicity by knockout of Sod1

APAP, also known as acetaminophen or paracetamol, is the active component of Tylenol and many other over-the-counter analgesics. A life-threatening hepatotoxicity of APAP overdose depends on the liver enzyme CYP2E1 (cytochrome P-450 2E1) that catalyzes biotransformation of APAP to a highly reactive intermediate, N-acetyl-p-benzoquinoneimine (NAPQI), which in turn can cause depletion of hepatic GSH and excessive liver necrosis (306). Interestingly, genetic deletion of several antioxidant enzymes yields increased APAP resistance in mice, which has been reported for Gstp1 (255), TrxR1 (see below), and Sod1.

While an intraperitoneal injection of 600 mg/kg APAP results in 75% mortality in wild-type mice within 20 h, all such treated Sod1−/− mice survive for the entire 70 h duration of study (379). Moreover, the Sod1−/− mice survived nearly three times as long as, and showed much less hepatic injuries than, wild-type mice following both higher (1,200 mg/kg) and lower (300 mg/kg) doses of APAP injection, respectively. As shown in Figure 1, this astonishing resistance to APAP intoxication is associated with at least four separate mechanisms. First, these mice have a 50% reduction in activity of the NAPQI-producing enzyme CYP2E1 in liver. The downregulated CYP2E1 activity thus helps attenuate NAPQI formation and the resultant GSH depletion and protein adduct formation. Indeed, hepatocytes isolated from Sod1−/−Gpx1−/− mice display a lower susceptibility to APAP-induced cell death, but higher susceptibility to NAPQI toxicity compared with cells from wild-type mice (754). Second, hepatic protein nitration plays a crucial role in mediating APAP-induced hepatotoxicity (343). Knockout of Sod1 nearly completely blocks APAP-induced hepatic protein nitration (379). This is intriguing as the enzyme knockout or depletion presumably elicits elevated O2 production and thus subsequent peroxynitrite formation for protein nitration, provided that NO is available. Strikingly, SOD1 was previously shown to catalyze peroxynitrite-mediated nitrotyrosine formation in vitro (298). Later, the enzyme was demonstrated to be required for the protein nitration mediated by APAP or LPS in murine liver (758). Third, compensatory inductions of other protective antioxidant enzymes (379, 756) and, fourth, a blunted cell death signaling (757) also attribute to the APAP resistance of the Sod1−/− mice. Seemingly, the above-described Sod1 deficiency-derived protection against the APAP overdose is cytosolic-specific. The mitochondrial Sod2+/− mice are actually more prone to the APAP-induced liver toxicity than their wild-type controls, potentially through prolonged JNK activation, exaggerated mitochondrial dysfunction with nuclear DNA fragmentation, and necrosis (200, 545). It remains unclear whether Sod2+/− mice are altered with expression of CYP2E1 and metabolism of APAP. However, SOD2 may serve a more important role than SOD1 as mitochondrion is a main target of the APAP toxicity. As the protein level of hepatic CYP2E1 in the Sod1−/− is not altered (379), the activity loss probably results from an oxidative modifcation. However, another group failed to detect similar decreases in the baseline activity of CYP2E1 in Sod1−/− mice, despite conflicting effects of ethanol on the enzyme activity between their own studies (133, 329).

FIGURE 1.

FIGURE 1.Protective mechanisms conferred by the Sod1 knockout against APAP toxicity. The associated mechanisms include the following: 1) inhibition of protein nitration; 2) upregulation of GR and TrxR activities; 3) downregulation of microsomal P-450 enzyme CYP2E1 activity, decreasing NAPQI production and the subsequent GSH depletion; and 4) inhibition of cell death signaling. Collectively, the Sod1−/− mice, as shown in the bottom graph, display a 100% survival rate over 70 h, while 75% of the wild-type mice die within 20 h after an intraperitoneal injection of 600 mg/kg body wt APAP. APAP, acetaminophen; Bcl-XL, B-cell lymphoma-extra large; CYP2E1, cytochrome P-450 2E1; GR, glutathione reductase; GSH, glutathione; IκBε, inhibitor of NFκB, epsilon; JNK, c-jun NH2-terminal kinase; NAPQI, N-acetyl-p-benzoquinoneimine; NFκB, nuclear factor kappa-light-chain-enhancer of activated B cells; p21, cyclin-dependent kinase interacting protein 1; PARP, poly(ADP-ribose) polymerase; and TrxR, thioredoxin reductase. [The survival data are modified from Lei et al. (379).].


2. Protection against irradiation-induced neuronal damages by knockout of Sod

Knockout of Sod1 or Sod2 decreases a baseline of neurogenesis, but ameliorates radiation-induced decline of neurogenesis (183, 286) (Figure 2). Following irradiation, Sod2+/− mice preserve normal hippocampal-dependent cognitive functions and normal differentiation patterns for newborn neurons and astroglia, which otherwise are damaged in irradiated wild-type mice. However, irradiation leads to a disproportional reduction in newborn neurons of the Sod2+/− mice following behavioral training, suggesting that Sod2 haploinsufficiency renders newborn neurons susceptible to metabolic stress (126). In contrast, irradiation of Sod3−/− mice enhances hippocampus-dependent cognition and decreases hippocampal nitrotyrosine formation (540). These results suggest that chronically elevated O2 anion levels and/or the lower production of H2O2 resulting from a Sod3 knockout may be protective against irradiation-induced damages in neurogenesis and cognition. In line with this result, overexpression of SOD3 impairs long-term learning and potentiation in hippocampal area CA1, further suggesting that O2, rather than being considered exclusively neurotoxic, may also be a signaling molecule necessary for normal neuronal function (644). The underlying molecular mechanisms and signaling pathways for these phenotypes await further investigation. Likewise, knockout of Sod1 enhances recovery after closed head injury-induced brain trauma in mice, which is associated with attenuated activation of NFκB and subsequent decreased death-promoting signals due to downregulated H2O2 production (41).

FIGURE 2.

FIGURE 2.Protections conferred by knockouts and haploid insufficiencies of Sod1, Sod2, and Sod3 against neural and cognitive damages induced by irradiation and brain trauma. Whereas mechanisms for the protections against the irradiation-induced damages await further investigation, the enhanced recovery from the brain trauma in the Sod1−/− mice is associated with attenuated H2O2 production and the subsequent NFκB activation. NFκB, nuclear factor kappa-light-chain-enhancer of activated B cells.


3. Neurological disorders associated with overexpression of SOD1/Sod1

SOD1 expression is associated with two types of neurological diseases: Down's syndrome (with elevated SOD activity) and ALS (associated with SOD1 mutations) (569, 612). Indeed, SOD1-overexpressing mice manifest certain abnormalities that resemble physiological effects seen in Down's syndrome, including withdrawal and destruction of some terminal axons and development of multiple small terminals (23, 24), a defect in platelet's dense granule responsible for the uptake and storage of blood serotonin (580), thymus and bone marrow abnormalities (524), and an impairment of hippocampal long-term potentiation (201). Meanwhile, SOD1 overexpression causes mitochondrial vacuolization, axonal degeneration, and premature motor neuron death as well as accelerates motor neuron degeneration in mice expressing an ALS-inducing SOD1 mutant (303). The SOD1 overexpression also impairs muscle function and leads to typical signs of muscular dystrophy in mice (525, 550). In fact, transgenic mice overexpressing SOD1 display aberrant protein expression profiles in neurons and mitochondria of hippocampus (605, 606), indicating that elevated SOD1 activity in Down's syndrome is not just a side effect or a compensation in response to the increased oxidative stress, but may be part of the cause for the pathophysiology.

Overexpression of SOD1 impairs peripheral nerve regeneration and increases development of neuropathic pain after sciatic nerve injury with a disturbed inflammatory reaction at the injury site (350), exacerbates abnormalities in hematopoiesis and radiosensitivity in a mouse model of ataxia-telangiectasia (529), and promotes aging as indexed by mitochondrial DNA deletion in the acoustic nerve of transgenic mice (120). Neurons from the SOD1 overexpressing mice exhibit higher susceptibility to kainic acid-mediated excitotoxicity, associated with a chronic pro-oxidant state as manifested by decreased cellular GSH and altered Ca2+ homeostasis (30). All these negative impacts, along with known biochemical and neurological mechanisms, of SOD1 overexpression on various neurological disorders are summarized in Figure 3.

FIGURE 3.

FIGURE 3.Induction or potentiation of various neurological disorders by SOD1 overexpression. While mice overexpressing SOD1 develop signs of Down's syndrome and muscular dystrophy, the overexpression promotes or exacerbates pathogenesis of other listed disorders including amyotrophic lateral sclerosis that is induced by the overexpression of mutant SOD1. Respective biochemical and neurological mechanisms for the impacts of SOD1 overexpression, along with their change directions (up and down arrows), are schematically shown for each of the disorders. GSH, glutathione.


In contrast, other studies have shown either negligible effects of Sod1 overexpression on toxicities induced by neurotoxins including kainite, glutamate, and N-methyl-d-aspartate (NMDA) (347, 729) or even protections against similar insults in vivo (260, 586) and in vitro (53, 92). These seemingly contradictory findings may be confounded in part with differences in extents of Sod1 overexpression, acute versus chronic experimental settings, the timing of observations, and the cellular capacity of H2O2 catabolism at the testing condition. For example, when treated with a O2 donor, overexpression of SOD1 increases neuronal vulnerability due to increased H2O2 accumulation, while overexpression of the gene in astrocytes that exhibit a greater H2O2 catabolism capacity than do neurons actually leads to an increased resistance to O2 toxicity (104). Therefore, the “paradoxical” function of SOD1/Sod1 overexpression in the central nervous system may largely rely on 1) whether the generated extra H2O2 results in a burden beyond affordable cellular clearing capacity, 2) whether the induced burst of O2 is more detrimental to cell survival than the converted extra amount of H2O2, and 3) whether effects of the enzyme expression are unrelated to its enzymatic activity.

4. Impaired immune functions and detrimental effects by overexpression of SOD1/Sod1

Overexpression of SOD1 in intraperitoneal macrophages decreases their microbicidal and fungicidal activity, along with increased intracellular production and release of H2O2, decreased extracellular release of O2, and inhibited NO production following endotoxin stimulation (456). It was intriguing why enzymatically derived NO production became decreased when O2 anion levels were diminished. The authors noted that nitrocompound metabolism in macrophages was affected by the overproduction of SOD1, but did not give mechanistic explanations. Possibly the reduced activities of NFκB and Erk1/2 in the SOD1 overexpressing macrophages, which are upstream regulators of iNOS, lead to downregulation of iNOS expression and thus lower NO production. However, this hypothesis remains to be experimentally confirmed. Transgenic mice overexpressing SOD1 show no increased resistance to TNF-α-induced endotoxic shock (144), but a higher sensitivity to malaria infection as reflected by an earlier onset and increased rate of mortality (220) as well as activation-induced DNA fragmentation in their splenic T cells (513).

Doubling the expression of SOD1 does not extend, but instead causes a slight reduction of lifespan in mice (284). Likely due to elevated chronic oxidative stress, Sod1 overexpression leads to an increased heart rate variability (646) and accelerates the loss of cone function (668). Contrary to its protection against most ischemic injuries, overexpression of SOD1 in utero during ischemic reperfusion injury led to brain damage in both adult and fetal mice (383). Sod1−/− mice exposed to chronic ethanol consumption exhibit decreased alcohol dehydrogenase activity and little induced CYP2E1 activity, which suppresses ethanol metabolism and precludes the resultant steatosis (133), while these mice are more susceptible to the acute ethanol-induced liver injury (329). This apparently contrasting impact of Sod1 knockout on injuries associated with either acute or chronic ethanol intake underscore the stress-type and/or temporal dependence of the function and (patho)physiological relevance of this enzyme.

5. Diverse effects of SOD2/Sod2 overexpression on alcohol intoxication and cancer cell survival

While overexpression of Sod2 protects against liver mitochondrial DNA depletion and respiratory complex dysfunction after alcohol binge exposure via inhibition of the formation of peroxynitrite (433), the overexpression aggravates prolonged (7 wk) alcohol intake-induced hepatic toxicity (368, 433) (Figure 4). The prolonged ethanol intake selectively triggered hepatic iron elevation, lipid peroxidation, respiratory complex I protein carbonyls and dysfunction, mitochondrial DNA lesion, and depletion in Sod2 overexpressing mice. Because administration of an iron chelator (deferoxamine) prevents all these adverse effects, hepatic iron accumulation is likely the crucial factor for the metabolic disorder (368). It has been suggested that alcohol administration decreases the expression of hepcidin, leading to abnormally active duodenal ferroportin and increased intestinal absorption of iron, which gradually increases hepatic iron accumulation (246). Although it remains unclear why in the referenced study (368) the iron overload was only found in Sod2 overexpressing mice, elevated Sod2 activity was linked to hepatic iron accumulation through modulation of iron homeostasis proteins in alcoholic patients (481, 633).

FIGURE 4.

FIGURE 4.Aggravation of prolonged alcohol intake-induced hepatic toxicity by Sod2 overexpression. The main proposed mechanism is that Sod2 overexpression leads to accumulation of hepatic iron that partially replaces manganese in the active site of Sod2 to form Fe-Sod2. Consequently, Fe-Sod2 catalyzes production of hydroxyl radicals from H2O2 that cause lipid peroxidation and mitochondrial damage. Likely, the increased hepatic iron and H2O2 may also enhance production of hydroxyl radicals through Fenton reactions. In fact, the exacerbated effect of Sod2 overexpression can be prevented by iron chelators. Meanwhile, the anticipated diminished levels of superoxide anion may remove its beneficial role in limiting the propagation of lipid peroxidation and blunt the ethanol induction of iNOS and subsequent upregulation of PGC-1, leading to mtDNA depletion. Fe-Sod2, iron-substituted superoxide dismutase 2; iNOS, inducible nitric oxide synthase; mtDNA, mitochondrial DNA; NO, nitric oxide; PGC-1, peroxisome proliferator activated receptor gamma coactivator 1.


A proposed mechanism for aggravated hepatotoxicity by Sod2 overexpression may be as follows: the hepatic iron overload could lead to a decreased mitochondrial manganese uptake and increased misincorporation of iron in the active site, forming Fe-substituted Sod2. The Fe-Sod2 is stable and lacks SOD activity, but gains hydroxyl radical generating activity in the presence of H2O2, which in turn is generated by the manganese-SOD. Consequently, increased production of hydroxyl radicals could lead to the above-mentioned lipid peroxidation and other oxidative injuries (368). Apparently, increased hepatic iron and H2O2 might also generate hydroxyl radicals through Fenton reactions. The anticipated diminished O2 anion levels due to Sod2 overexpression might furthermore remove its beneficial roles in limiting propagation of lipid peroxidation and blunt alcohol-induced increases of inducible NO synthase and subsequent upregulation of peroxisome proliferator activated receptor gamma coactivator 1 (PGC-1), which otherwise promotes mitochondrial DNA replication (368). Another contributing factor could be the decrease in the mitochondrial transcription factor A (Tfam) in Sod2 overexpressing mice following alcohol administration. However, further studies are required to clarify the different cause-effect relationships with regards to the observed phenotypes. It should be noted that in rats, overexpression of Sod2 in liver prevents steatosis, inflammation, necrosis, and apoptosis following prolonged alcohol (4 wk) administration (694). Thus there may also be different impacts between species of Sod2 overexpression on ethanol metabolism and intoxication.

Recently, differential roles for SOD2 have been proposed between early and late stages of carcinogenesis. At the early stage, a lower SOD2 level may facilitate transformed phenotypes by potentiating mitochondrial defects, whereas at the later stage a higher SOD2 level protects cell from mitochondrial injury and contributes to tumor growth and metastasis (149). The roles of SOD2 become further complicated when cancer cells are challenged with increased oxidative stress. Overexpression of SOD2 in HeLa cervical cancer cells promotes their growth when growth factors are withdrawn, suggesting that SOD2 may promote tumor-cell survival in vivo at conditions unfavorable to cell growth by counteracting the intracellular oxidative processes that can additively impair cell growth and viability (514). Moreover, overexpression of SOD2 promotes survival of cancer cells treated with radiation, cytokines, or drugs (263, 387, 432, 469, 632), likely through activation of NFκB and AP-1 signaling by the SOD2-mediated conversion of H2O2. Therefore, overexpression of SOD2 may promote cancer due to increased cancer cell resistance to the cytotoxicity of therapeutic treatments.

B. Catalase

1. Diabetic developments induced by catalase overexpression

The β cell-specific overexpression of rat Cat in nondiabetic background mice shows no detrimental effects on islet function (717) and protects against the diabetogenic effect of STZ (99, 717). However, this type of overexpression provides no protection against cytokine-mediated toxicity in isolated islets, despite a suppression of ROS formation (99, 717). Interestingly, overexpression of CAT in mitochondria, compared with that in cytoplasm, confers stronger protections against the cytokine-induced cytotoxicity in insulin-producing cells (230, 407), indicating an important role of mitochondrial ROS in the cytokine toxicity of autoimmune diabetes.

Strikingly, β cell-specific overexpression of Cat in nonobese diabetic mice accelerates spontaneous diabetes onset in males and cyclophosphamide-induced diabetes in both males and females, and sensitizes isolated islets to cytokine injuries. Figure 5 depicts several described divergent effects of catalase overexpression on susceptibilities to diabetes, but none of these effects is fully understood mechanistically. There was a downregulation of Akt/Foxo1/Pdx1 survival pathway in islets associated with the cyclophosphamide-induced autoimmune type 1 diabetes (390). It was suggested that insulin/insulin-like growth factor I (IGF-I)-mediated phosphorylation of Akt might be downregulated by PTP-1B (a tyrosine phosphatase) that is inhibited by ROS (H2O2) and catalase overexpression prevented the ROS inhibition of PTP-1B (390, 574). Although there are no direct experimental data to support these notions, maintaining adequate intracellular H2O2 may be needed for activating protective responses of β cells in autoimmune type 1 diabetes.

FIGURE 5.

FIGURE 5.Divergent effects of catalase overexpression on susceptibility to diabetes. The β-cell specific overexpression of catalase in nonobese diabetic mice leads to early onset of spontaneous diabetes in male mice, with accelerated occurrence of diabetes by inhibition of the Akt-Foxo1-Pdx1 survival pathway in islets following cyclophosphamide administration. In contrast, such overexpression prevents diabetogenic effects of streptozotocin in nondiabetic mice. In insulin-producing cells, overexpression of catalase in mitochondria renders strong resistances to cytokine-induced cytotoxicity, whereas its overexpression in cytoplasm leads to weak protection. Akt, protein kinase B; Foxo1, forkhead box protein O1; Pdx1, pancreatic and duodenal homeobox 1.


In contrast, Cat overexpression consistently protects against diabetic nephropathy (58, 603) or insulin resistance-induced cardiac contractile dysfunctions (160). Overexpression of Cat also attenuates high glucose-induced reduction of endothelial cell tight-junction proteins and the subsequent brain blood barrier (BBB) dysfunction in diabetes (402). These data suggest differential roles of catalase in pancreas and other organs in diabetic versus physiological conditions.

2. Cell type-dependent inhibition of proliferation by catalase overexpression

Elevating catalase activity, similar to that of Sod, may alter the sensitivity of cancer cells to chemotherapy (216, 416). Overexpressing CAT in the cytosolic and especially in the mitochondrial compartments of HepG2 cells potentiates TNF-α-induced apoptosis by promoting activation of caspase-3 and -8 (26). In contrast, overexpression of Cat in a murine lymphoid cell line enhances resistance to dexamethasone-induced apoptosis and exhibits increased net tumor growth in nude mice, which is associated with a delay of mitochondrial cytochrome c release and altered glucose and energy metabolism (648, 649). Overexpression of CAT inhibits proliferation of endothelial cells (740) and vascular smooth muscle cells (66, 602) by suppressing Erk1/2 and p38 MAPK signaling (602) and promoting a Cox2-dependent apoptosis (66). This highlights the need for a physiological level of endogenous H2O2 for survival and proliferation of vascular cells. Interestingly, the proliferation rate is elevated in vascular smooth muscle cells of Sod1+/− and Sod2+/− mice, along with higher activity of divergent mitogenic signaling pathways. The heterozygosity of Sod1 leads to preferential activation of Erk1/2 and p38 MAPK, while that of Sod2 causes activation of JAK/STAT pathway in smooth muscle cells (422). This opposite outcome is intriguing, because overexpression of Cat presumably diminishes intracellular H2O2 whose formation would be supposed to be lower due to the Sod haplodeficiency. Nevertheless, these diverse effects underscore the physiological importance to tightly regulate intracellular H2O2 levels for control of vascular cell proliferation. Furthermore, specific overexpression of CAT in myeloid lineage cells impairs perfusion recovery associated with fewer neovascularization and blunted inflammatory response following a femoral artery ligation, suggesting that H2O2 derived from myeloid cells such as macrophages plays a key role in promoting neovascularization in response to ischemia and in the development of ischemia-induced inflammation (269). Notably, decreases of H2O2 levels upon overexpression of catalase were verified by direct assays in vascular smooth muscle cells and myeloid cells in the above-mentioned studies, but only by indirect methods in endothelial cells and lymphoid cells.

C. GPX Family

1. Improved insulin sensitivity and decreased insulin synthesis upon knockouts of Gpx1 and Sod1

While knockouts of Gpx1 and Sod1 impair islet function, pancreas integrity, and body glucose homeostasis, these mice present improved insulin sensitivity in liver and muscle (680, 684). This improvement is mainly associated with an increased phosphorylation of muscle Akt at Thr308 and Ser473 after injection of insulin (684) (Figure 6). Presumably, this “unanticipated” benefit is attributed to elevated intracellular ROS that inhibit protein phosphatase activities and thereby attenuate dephosphorylation of Akt (33, 680). Moreover, an increased IRβ protein in the liver of the Sod1−/−, but not in the Gpx1−/−, mice may also contribute to the improvement (684). Meanwhile, Gpx1−/− mice are resistant to the high-fat diet-induced insulin resistance and show favorable responses including decreased expression of gluconeogenic genes (G6pc, Pck1 and Fp1), increased glucose uptake by white gastric and diaphragm skeletal muscles through membrane docking of glucose transporter 4 upon AS160 phosphorylation on Thr642, and enhanced insulin-induced oxidation of phosphatase and tensin homolog (Pten) and PI3K/Akt signaling (406) in their embryonic fibroblast cells.

FIGURE 6.

FIGURE 6.Comparative mechanisms for improved insulin sensitivity in Sod1−/− and Gpx1−/− mice. Knockout of Sod1 elevates hepatic IRβ protein and muscle Akt phosphorylation after insulin stimulation, whereas knockout of Gpx1 induces only the latter. Meanwhile, embryonic fibroblast cells from Gpx1−/− mice are manifested with enhanced Pten oxidation and PI3K/Akt activation after insulin addition. These changes are presumably (dashed arrows) upstream of Akt phosphorylation and result in improved insulin sensitivity. However, such impact of Sod1 deletion has not been tested yet (question mark). In addition, Gpx1−/− mice fed a high-fat diet display, following insulin challenge, enhanced glucose update through membrane docking of glucose transporter 4 upon AS160 phosphorylation on Thr642. Akt, protein kinase B; AS160, the 160 kDa substrate of Akt; IRβ, β subunit of insulin receptor; PI3K, phosphatidylinositol-3-kinase; Pten, phosphatase and tensin homolog.


Comparatively, the Sod1 knockout exerts stronger impacts on insulin synthesis and secretion, glucose and lipid metabolism, and islet integrity than that of Gpx1 (684). Simultaneous ablation of both enzymes does not result in additive or more severe metabolic outcomes. The Sod1−/− mice show more apparent pancreatitis than the Gpx1−/− mice that are more susceptible to the cerulein-induced amylase increase. Although hypoinsulinemia and decreased pancreatic β cell mass are caused by knockouts of both of Gpx1 and Sod1 via downregulation of the key transcription factor Pdx1 in pancreatic islets, the former seems to decrease only Pdx1 protein, whereas the latter exerts suppressions at three levels of the Pdx1 regulation: epigenetic, mRNA, and protein (684) (Figure 7). Likewise, knockout of Sod1, but not Gpx1, upregulates protein phosphatase 2b/sterol responsive element binding protein (Srebp)-mediated lipogenesis and downregulates the Ampk-mediated gluconeogenesis (680). Apparently, there are several overlapping as well as distinctive mechanisms for Sod1 and Gpx1 in regulation of glucose homeostasis and lipid metabolism (378). It should also be noted that reductive stress may be as destructive as oxidative stress in the etiology of diabetes and obesity (753).

FIGURE 7.

FIGURE 7.Distinctive mechanisms between knockouts of Sod1 and Gpx1 in lowering pancreatic islet β cell mass and plasma insulin concentration via downregulation of the key transcription factor Pdx1. While knockout of Gpx1 decreases only the Pdx1 protein in islets, knockout of Sod1 exerts suppression at three levels of Pdx1 regulation: epigenetic, mRNA, and protein. The downregulation of Pdx1 mRNA and protein upon Sod1 knockout coincides with decreased mRNA and protein levels of Foxa2, a transactivator of Pdx1, as well as attenuated binding of Foxa2, H3 acetylation, and H3K4, trimethylation in the proximal region of the Pdx1 promoter. Foxa2, forkhead box A2; H3, histone-3; K4, lysine-4; ORF, open reading frame; Pdx1, pancreatic and duodenal homeobox 1.


2. Potentiation of peroxynitrite-induced toxicity by GPX1/Gpx1

Peroxynitrite (PN) represents a major RNS formed from reaction of O2 with NO, which occurs at a diffusion-limited rate (39). Although PN induces nitration of a variety of biomolecules, a major activity indicator is the nitrosylation of protein tyrosine residues (39). Peroxynitrite-mediated protein nitration is indeed involved in the pathogenesis of many human diseases (297, 530). Impacts and mechanisms of the influence of GPX1 activity on peroxynitrite-induced oxidative damage have been studied in different systems (198, 610). Figure 8 summarizes “paradoxical” roles and mechanisms of bovine GPX1, Gpx1 knockout, and GPX1 overexpression in coping with the PN-mediated protein nitration and toxicity in these systems.

FIGURE 8.

FIGURE 8.Paradoxical roles of bovine GPX1, Gpx1 knockout, and GPX1 overexpression in coping with PN-mediated protein nitration and toxicity in cell-free system, primary hepatocytes, and mice. Different insult-mediated responses with net impacts of enzyme expression, and reported or proposed mechanisms are summarized in this figure. APAP, acetaminophen; DQ, diquat; GSH, glutathione; GST, glutathione-S-transferase; PN, peroxynitrite; SNAP, S-nitroso-N-acetyl-penicillamine.


Using a cell-free system, Sies et al. (610) found that GPX1 can serve as a peroxynitrite reductase. However, that function of GPX1 could not be verified by Fu et al. (198) using primary hepatocytes isolated from Gpx1−/− mice. In stark contrast, Gpx1−/− hepatocytes are instead extremely resistant to peroxynitrite-induced DNA fragmentation, cytochrome c release and caspase-3 activation, GSH depletion, protein nitration, and cell death (198). Interestingly, treating hepatocytes with S-nitroso-N-acetyl-penicillamine (SNAP; a NO donor) in addition to diquat (O2/H2O2 donor) produces synergistic cytotoxicity, and protein nitration induced by these two pro-oxidants together is attenuated in Gpx1−/− cells (197). While knockout of Gpx1 in mice exerts partial protection on the APAP- or LPS-induced hepatic toxicity and protein nitration (343, 755, 756), overexpressing GPX1 sensitizes mice to the APAP-induced hepatotoxicity and lethality (458). The metabolism of APAP in GPX1 overexpressing mice leads to a substantial decrease in the replenishment of GSH in liver and blood compared with the controls. In contrast, overexpressing GPX3 and Sod1 in the same study renders mice resistant to the APAP toxicity. These observations again underscore the complexity or unpredictability of seemingly similar antioxidant enzymes in coping with a given oxidative insult.

3. Protection against kainic acid-induced lethality and seizure by Gpx1 knockout

Kainic acid is an analog of glutamate that is widely used to induce limbic seizures and model the disease of epilepsy in rodents (40, 217). Administration of the compound activates NMDA receptors in hippocampus and other vulnerable brain regions (31, 43). As an event following NMDA activation (364, 365), there is increased oxidative stress including formations of O2, NO, and PN in the central nervous system after kainic acid injection (217, 452, 568). Thus antioxidants such as ascorbate, GSH, and EUK-134 (a synthetic SOD and catalase mimic) can decrease the neurotoxic effects of kainic acid and its seizure-associated neuropathology (421, 575). Strikingly, Gpx1−/− mice are much more resistant to kainic acid-induced seizure (frequency and interval), neuronal injury, and lethality compared with wild-type controls (309). This increased resistance involves inactivation of the NMDA receptor via thiol oxidation of its NMDA receptor-1 subunit, possibly due to elevated H2O2 levels in the brain of Gpx1−/− mice, and subsequent attenuation or block of the kainic acid-induced oxidative injuries (309) (Figure 9). As described above, neurons from SOD1 overexpressing mice exhibit elevated susceptibility to the kainic acid-mediated excitotoxicity (30). Therefore, certain levels of ROS or chronic oxidation in the brain are needed for a functional NMDA receptor, with long-term use of antioxidants possibly thereby leading to detrimental rather than protective effects.

FIGURE 9.

FIGURE 9.Mechanisms of protection conferred by Gpx1 knockout against kainic acid-induced neurotoxicity. Gpx1 deficiency may elevate H2O2 production that can oxidize thiols in the NMDA receptor-1 subunit, which deactivates the NMDA receptor and subsequently attenuates or blocks kainic acid-induced oxidative stress and injuries. This oxidative stress can also be protected by antioxidants. EUK-134, a synthetic SOD and catalase mimic; GSH, glutathione; NMDA, N-methyl-d-aspartate.


4. Type 2 diabetes-like phenotypes induced by Gpx1 overexpression

Global overexpression of Gpx1 in nonobese or nondiabetic mice results in hyperinsulinemia, hyperglycemia, hyperlipidemia, insulin resistance, β-cell hypertrophy, and obesity at 6 mo of age (445). Diet restriction can prevent all these phenotypes except for hyperinsulinemia and hypersecretion of insulin after glucose-stimulation (686). Thus these two phenotypes represent primary effects of Gpx1 overproduction and seem to be mediated by upregulation of a key transcription factor (Pdx1) for β-cell differentiation and insulin synthesis and secretion, as well as downregulation of the insulin secretion inhibitor mitochondrial uncoupling protein 2 (Ucp2). The insulin resistance in these Gpx1 overexpressing mice may be attributed to less oxidative inhibition of protein tyrosine phosphatases due to diminished intracellular ROS (H2O2) levels upon higher Gpx1 activity, leading to accelerated dephosphorylation of IRβ and Akt after insulin stimulation (445, 686). Meanwhile, Gpx1 overexpression also affects transcripts, proteins, and functions of other pro-insulin genes, lipogenesis rate-limiting enzyme genes, and key glycolysis (GK) and gluconeogenesis (PEPCK) enzymes in islets, liver, and muscle (526, 721). Figure 10 highlights the major pathways and modes of action in relation to insulin production and insulin responses, illustrating how Gpx1 overexpression can induce type 2 diabetes-like phenotypes. Dietary selenium deficiency precludes Gpx1 overproduction in these mice and partially rescues their metabolic syndromes by modulating or reversing these molecular and biochemical changes (721). Similarly, dietary selenium levels have indeed been shown to affect glucose metabolism and insulin sensitivity (362). However, β-cell-specific overexpression of GPX1 in db/db mice with mutated leptin receptor rescues β-cell dysfunction with reversed signs of diabetes at 20 wk of age (229, 244). It should be noted that islets have relatively low baseline Gpx1 activity but display one of the highest overproductions of Gpx1 activity among all tissues in the global Gpx1 overexpressing mice. Collectively, it seems clear that GPX1/Gpx1 overproduction is beneficial at diabetic or obese pathophysiological conditions, but becomes deleterious if triggered in healthy mice with normal metabolic status.

FIGURE 10.

FIGURE 10.Molecular and biochemical mechanisms for the type 2 diabetes-like phenotypes induced by Gpx1 overexpression in mice. The diminished H2O2 accumulation in pancreatic islets may enhance β cell mass and insulin synthesis and secretion via modulation of key signaling genes and proteins at epigenetic, mRNA, and/or protein levels. These effects lead to hyperinsulinemia and hypersecretion of insulin. Meanwhile, Gpx1 overexpression also impairs insulin responsiveness in liver and muscle and disturbs lipogenesis, glycolysis, and gluconeogenesis in those tissues. The reported modes of action for those impacts include modulation of key gene expression, protein function, and enzyme activities. The outcomes from these effects in insulin-responsive tissues are reflected by insulin resistance, hyperglycemia, hyperlipidemia, and obesity. The overall phenotypes from GPx1 overexpression in either insulin-producing or insulin-responsive tissues resemble type 2 diabetes. Representative key factors for each of the main pathways or phenotypes are listed in brackets. Acc1, acetyl-coenzyme A carboxylase 1; Beta2, neurogenic differentiation 1; Cat, catalase; Cfos, fbj murine osteosarcoma viral oncogene homolog; Fasn, fatty acid synthase; Foxa2, forkhead box a2; Ins1, insulin 1; IRβ, the β-subunit of insulin receptor; Kir6.2, the KCNJ11 subunit of ATP-sensitive K+ channel; p53, transformation related protein 53; Pdx1, pancreatic and duodenal homeobox 1; Pparγ, peroxisome proliferator-activated receptor γ; Pregluc, preproglucagon; Sur1, sulfonylurea receptor; Ucp2, uncoupling protein 2; Akt, protein kinase B; GK, glucokinase; PEPCK, phosphoenolpyruvate carboxykinase; Δψ, mitochondrial membrane potential.


5. Intriguing roles of GPX enzymes in carcinogenesis

A number of studies have revealed cancer type-, stage-, and tissue-dependent impacts of GPX enzymes on carcinogenesis (Figure 11). Global GPX1 overexpression sensitizes mice to skin tumor formation induced by 7,12-dimethylbenz(a)anthracene/12-O-tetradecanoylphorbol-13-acetate (DMBA/TPA) (413), whereas adenoviral delivery of GPX1 to pancreatic tumor xenografts actually suppresses the tumor growth in nude mice (403). While mechanisms for differential roles of GPX1 between skin and pancreatic tumors remains elusive, a deficiency of Gpx1 in cancer-free naked mole rats (323) suggests the enzyme to be dispensable for cancer prevention at least in this particular species. As discussed in section II, chronic colitis (174) and inflammation-driven intestinal cancer (118) are observed in Gpx1−/−Gpx2−/− mice, but not in Gpx2−/− mice unless additional stress is employed (678). This implies dose-dependent or overlapping roles of the two Gpx enzymes in this regard.

FIGURE 11.

FIGURE 11.Intriguing roles of GPX isoenzymes in carcinogenesis. A: overexpression of GPX1 in mice promotes DMBA/TPA-induced skin cancer, whereas adenoviral delivery of GPX1 to pancreatic tumor xenografts slows tumor growth in nude mice. This contrast illustrates tissue- or stage-specific roles of GPX1 in carcinogenesis. As depicted in the middle, Nrf2 and β-catenin that are associated with cancer were shown to upregulate GPX2 expression in cultured human cells. Knockout of Gpx2 either stimulates or inhibits AOM-DSS-induced intestinal tumorigenesis at early or late stages, respectively. The relative stage-specific effects are indicated on the pink-colored triangle box, exemplifying the temporal dependence of the GPX enzyme in carcinogenesis. In addition, knockout of Gpx3 in mice promotes AOM/DSS-induced colitis-associated carcinoma, but knockdown of the enzyme by shRNA inhibits leukemia stem cell renewal. This comparison illustrates the cancer type- and/or model-specific role of the GPX enzyme in carcinogenesis. AOM/DSS, azoxymethane/dextran sodium sulfate; DMBA/TPA, 7,12-dimethylbenz[a]anthracene/12-O-tetradecanoylphorbol-13-acetate; Nrf2, NF-E2-related factor 2; shRNA, small hairpin RNA.


Roles of GPX2 in carcinogenesis vary with cellular metabolic contexts (61, 473). In healthy (normal or precancerous) cells, the enzyme helps maintain self-renewing of the gastrointestinal epithelium, suppress inflammatory processes, and thereby inhibit carcinogenesis (61). Loss of Gpx2 induces apoptosis, mitosis, and elevated Gpx1 expression in the intestine of mice (62, 186). In cancerous cells, however, the anti-apoptosis function of the enzyme may promote their growth and migration. After being treated with intestinal carcinogens azoxymethane/dextran sulfate, Gpx2−/− mice fed a selenium-inadequate diet (0.08 mg Se/kg) showed increased tumor numbers but decreased sizes, compared with the wild-type controls (352). The elevated tumor numbers may reflect Gpx2-derived protection against carcinogenesis at the early stage of tumor formation, whereas the declined tumor sizes imply promotion of Gpx2 on tumor growth during the late stages of tumorigenesis. In addition, upregulation of GPX2 in colorectal cancer (59, 187) and activation by cancer-associated NRF2 and Wnt pathways (28, 336) further suggest that the enzyme may indeed be involved in the pathophysiological processes.

Several studies have collectively demonstrated both positive and negative impact of GPX3 on carcinogenesis (62). While knockout of Gpx3 sensitizes mice to chemical-induced, colitis-associated carcinoma (32), knockdown of the gene by shRNA in leukemia stem cells decreased their competitiveness and self-renewal capability (256). Apparently, most of the findings on the roles of GPX enzymes in carcinogenesis await full elucidation of the molecular and cell biology mechanisms.

6. Paradoxical effects of TrxR1 targeting by genetic modulation or drug treatment

Although TrxR1 is an essential enzyme for mouse embryogenesis, the enzyme can be conditionally deleted in a wide range of differentiated tissues without apparent phenotype (Table 8, see above) and fully inhibited for at least a week in mice by gold compound treatment without overt toxicity (615). Strikingly, genetic deletion or full inhibition of TrxR1 can instead protect cells and tissues from oxidative challenges. As in the case of global knockout of Sod1 (see above), liver-specific Txnrd1−/− mice become highly resistant to APAP-induced hepatotoxicity (302, 520). It was also found that the TrxR1 enzyme, together with GSH, is a prime target for inhibition by NAPQI, which should help explain why APAP-derived NADPQI becomes more toxic than what is seen upon mere GSH depletion using inhibition of GSH synthesis (302, 520). The protective effects of TrxR1 deletion against APAP challenge are likely to be explained by compensatory upregulation of many Nrf2 targets in mice with hepatocytes lacking Txnrd1 to yield more robust GSH biosynthesis, glutathionylation, and glucuronidation systems following APAP overdose (302, 520). Indeed, “priming” of tissues for oxidative injuries by prior inhibition of TrxR1 has also been shown in lung tissue, where inhibition of the enzyme leads to better resistance to hyperoxia (63, 405), in most or all of these cases presumed to involve activation of the cell protective Nrf2 pathway (89).

Similar paradoxical roles of TrxR1 exist in carcinogenesis. Although liver-specific Txnrd1−/− mice were reported to display a much greater tumor incidence (90 vs. 16%) compared with wild-type mice after diethylnitrosamine induction (79), the Trx/TrxR1 system has also been found to promote tumor growth (19, 251). Tumors arising in mice after injection of Txnrd1-knockdown Lewis lung carcinoma (LLC1) cells are much smaller in size than those from mice injected with the control, malignant cells, and most importantly, these knockdown LLC1 cells lose their targeting construct or show attenuated metastasis (251, 730). The mechanisms by which the enzyme can either be cancer preventive or promoting cancer progression are not fully understood, but are likely to relate to different stages of carcinogenesis and perhaps also differ between cancer types. It is known that overexpression of TrxR1 in cancer cells correlates with tumorigenic properties, and downregulation of the enzyme inhibits growth of human hepatocarcinoma cells (204). As mentioned above, knockdown of TrxR1 in lung carcinoma cells reverses their tumorigenicity and invasive potential in a xenograft model (730). Therefore, TrxR1 enzymes have been suggested as potential targets for development of anticancer drugs (410, 485, 666). As loss of Txnrd1 renders tumors highly susceptible to pharmacological GSH deprivation, a concomitant inhibition of both GSH and TxrR systems was recently proposed to be a strategy to kill tumor cells (245, 430). In this context it should be noted that drug-targeted inhibition may not only inhibit TrxR1, but can also convert the enzyme to a pro-oxidant NADPH oxidase upon selective modification of its Sec residue (13).

We conclude that TrxR1 can exert “paradoxical” effects in three separate forms of the enzyme, i.e., no matter whether it is overexpressed, knocked down, or targeted by low-molecular-weight inhibiting compounds, either beneficial or detrimental physiological effects can be triggered depending on cellular context. This is summarized in Figure 12 and its diverse effects should be considered in studies aimed at understanding the physiological roles of this enzyme.

FIGURE 12.

FIGURE 12.Paradoxical effects of TrxR1 overexpression, genetic loss, or drug inhibition. Three separate states of TrxR1 can have either beneficial (top) or detrimental (bottom) effects in mammals, as summarized in this figure. Native TrxR1 promotes cell viability through diverse functions of the Trx system, including support of Prxs, Msrs, and RNR as well as modulation of redox signaling pathways. However, cancer cells may also rely on TrxR1 activity to proliferate and the enzyme can thus promote cancer progress as well as metastases. Genetic targeting of TrxR1 is embryonically lethal, while conditional knockout in differentiated tissues such as hepatocytes result in Nrf2 activation and an increased resistance to oxidative challenges. When targeted by inhibitors, the TrxR1 enzyme can also gain an NADPH oxidase activity in addition to loss of its native Trx1 reducing capacity, which further activates Nrf2 but can also trigger cancer cell death, toxic side effects in normal tissues, and increased dependence on GSH for survival. APAP, acetaminophen; Msrs, methionine sulfoxide reductases; NAPQI, N-acetyl-p-benzoquinoneimine; Nrf2, NF-E2-related factor 2; Prxs, peroxiredoxins; RNR, ribonucleotide reductase; Trx, thioredoxin; TrxR1, thioredoxin reductase-1.


E. Hazard of Trx Overexpression and Benefit of Grx Knockout

Increased Trx1 potentiates cadmium toxicity (218), whereas ablation of Grx1 renders mice resistant to the LPS-induced inflammation and macrophage activation associated with enhanced S-glutathionylation (4). The latter also enhances resolution of airway hyperresponsiveness and mucus metaplasia in allergic mice (270). Because the gene knockout also attenuates inflammation and expression of proinflammatory mediators in the lung, S-gluathionylation of specific target proteins may be beneficial to attenuate airway hyperrepsonsiveness like in asthma. Thus inhibitors of Grx1 may be of interest clinically. Plasma Grx1 concentration is increased in patients with diabetes (163). This may be linked to defective revascularization in diabetes, since Grx1 overexpressing mice have elevated soluble vascular endothelial growth factor receptor 1 and attenuated postischemia limb revascularization (478).

IV. MECHANISMS AND METABOLIC RELEVANCE

The apparently paradoxical outcomes in several cases of antioxidant enzyme overexpression or genetic deletion studies clearly challenge the “prevailing” view that these enzymes are only beneficial, or that ROS/RNS are solely toxic byproducts of aerobic metabolism. It is clear that controlled production of ROS/RNS is important in signaling and that under certain conditions, antioxidant enzymes exhibit pro-oxidant activities. In all aspects of redox biology, spatial-, tissue-, and temporal-specific dependences are crucial, which will also have an impact on the physiological functions of antioxidant enzymes (307, 465, 656).

The exact mechanisms for “paradoxical” outcomes of antioxidant enzyme knockout or overexpression should undoubtedly derive from the interplay of three factors: 1) the properties and roles of their ROS/RNS substrates and products, 2) the activities and functions of the antioxidant enzymes, and 3) the metabolic contexts in which these entities interact. Accordingly, we will here discuss a series of chemical, molecular, biochemical, and physiological mechanisms that need to be considered and that may help to explain the observed paradoxical roles of antioxidant enzymes. Contributions of reductant substrates such as GSH to the paradox are discussed in the context of antioxidant enzyme catalysis.

A. Multifaced Chemical Reactivity and Metabolic Roles of ROS/RNS

1. Dose-dependent impacts of ROS/RNS

Whereas excessive levels of ROS and RNS trigger oxidative stress, appropriate levels of ROS/RNS are required for redox signaling. Apparently, antioxidant enzymes are needed to suppress excessive production of ROS and RNS. Under certain conditions, however, insufficient ROS/RNS or elevated cellular reductants can be detrimental, or, conversely, elevated ROS/RNS may be beneficial. This explains in part dose-dependent effects of ROS/RNS or roles of their metabolizing enzymes. Transgenic mice with 2- to 3-fold increased Sod2 activity in major organs are phenotypically normal and fertile (542), while a higher overexpression of the enzyme to 2.5- to 8.7-fold activity above normal decreases body size and female fertility, and causes male infertility. Transgenic lines overexpressing 60- or 100-fold catalase activity are more resistant to doxorubicin-induced cardiac injury, but further overexpression to 200-fold or higher fails to provide protection (319). While the precise molecular explanations to these observations are unknown, they likely involve effects of site-specific localization, reactivity, steady-state levels of H2O2, as well as differential induction of compensatory pathways, as discussed below.

2. Detrimental effects of insufficient peroxides on redox signaling

Of the primary ROS, H2O2 is perhaps the most important for signaling (560), with both O2 and hydroxyl radicals having limited half-life and reactivity profiles unsuitable for diffusible signals (135, 192, 193). H2O2 is an ideal signaling agent because of its relatively long lifetime and selectivity for targeting of particular protein microenvironments (135, 181, 192, 193, 700). It can oxidize thiol groups of specific Cys residues to disulfides (S-S), sulfenic (S-OH), sulfinic (SO2H), and sulfonic (SO3H) acids (404). Overoxidation to sulfonic acid is not implicated in redox signaling, but contributes to oxidative stress due to its irreversibility. Peroxide-sensing proteins that utilize uniquely reactive Cys residues may include transcription factors (20, 487), kinases (651), phosphatases (451), ion channels (521), ubiquitin and small ubiquitin-related modifier (SUMO)-conjugating enzymes, ligases and adapter proteins (55, 157, 450, 743), as well as various metabolic enzymes (466). Most proteins to react with H2O2 are peroxidases, such as GPXs or PRXs, which in turn may propagate the oxidative signal to specific downstream targets in cells (698, 699).

Many signal transduction pathways are hard-wired to redox signaling networks, due to the large number of kinases and phosphatases having reactive Cys residues that affect their activities (70, 192, 193, 673). Deliberately produced peroxides can oxidize catalytic Cys residues in various protein tyrosine phosphatases (PTPs), thereby inactivating them (34, 146, 577, 650, 670). This in turn serves to enhance activation of related signaling pathways by preventing the PTP-catalyzed dephosphorylation of specific phosphorylated tyrosine residues. Apparently, this type of inhibition can be removed by a hyperactivity of peroxide-scavenging enzymes. As in the case of Gpx1-overexpressing mice that develop type 2 diabetes-like phenotypes (445, 686), the overproduced Gpx1 diminishes intracellular ROS production, reverts the inhibition of PTPs, accelerates dephosphorylation of IR and Akt after insulin stimulation, and thereby leads to insulin resistance. In contrast, knockout of Gpx1 causes accumulation of intracellular peroxide, which, via the same pathways, improves insulin sensitivity and renders the mice resistant to high-fat diet-induced insulin resistance (406, 684). However, the specific dose, temporal dynamics, and protein phosphatases that are targeted by H2O2 in redox signaling remain largely unclear.

3. Mixed effects of peroxynitrite in cell signaling

Peroxynitrite-mediated signaling pathways are not as firmly established as those involving H2O2. Traditional “antioxidant” ROS-scavenging enzymes like SOD1 and GPXs have been implicated in PN metabolism and are supposed to affect the related signaling pathways. PN can upregulate Src tyrosine kinases, the Akt pathway, and various mitogen-activated kinases (512). Because many mitogen-activated kinases like p38 and c-Jun are implicated in proapoptotic pathways, PN is considered to be a pro-death signaling molecule. In addition, PN has also been implicated in hypoxic signaling. Under hypoxia, cytochrome c oxidase exhibits nitrite reductase activity, reducing nitrite to NO instead of oxygen to water (85, 86). This NO then reacts with O2 to form PN, which can oxidize certain unknown protein targets that signal adaptation to hypoxia (158, 534). As discussed in greater detail below, SOD1 can either increase or decrease PN via its ability to control O2 fluxes. Thus paradoxical outcomes of Sod1 knockout or SOD1 overexpression may be in part derived from the unpredictable consequences of PN modulation (298, 512, 758). The same may also be true for the case of GPX1 (197–199), but the precise roles and mechanisms of SOD1 and GPX1 in regulating PN-mediated signaling are unclear.

B. Paradox-Related Properties of Antioxidant Enzymes

1. Prooxidant catalysis

Despite their well-known ROS/RNS scavenging capacity, some antioxidant enzymes may also promote oxidative/nitrosative stress. One example is the conversion of TrxR1 to a prooxidant enzyme upon targeting of its Sec residue by inhibitors, as discussed above. Another illustrative example relates to the peroxidase activity exhibited by SOD1 (398, 399, 551, 745). The peroxidase cycle of SOD1 involves peroxide reducing the Cu(II) center to form O2 radical and Cu(I), followed by another molecule of peroxide reoxidizing Cu(I) to form Cu(II) and hydroxyl radical. These reactions can occur under severe peroxide stress, with the resulting hydroxyl radicals subsequently being able to irreversibly oxidize metal coordinating His residues and thereby inactivate SOD1 (728). In the presence of carbonate, hydroxyl radicals can also oxidize carbonate to form carbonate radicals, which can in turn oxidize a variety of other substrates, including azide, urate, and nitrite (448, 745). However, it remains unclear to what extent this chemistry happens in vivo, and how much this contributes to the SOD1 toxicity.

In some circumstances, SOD1 can also promote aberrant protein nitration, either by enhancing PN production or by directly activating it for tyrosine nitration. Beckman and colleagues demonstrated that human SOD1 mutants that are zinc deficient, either due to mutations associated with ALS or by other interventions that limit zinc to the protein, are better at catalyzing the reduction of dioxygen to O2, thus providing a pool of O2 that can react with nitric oxide to form PN (173, 655). The PN can then go on to nitrosylate and irreversibly damage various biomolecules, serving as another mechanistic basis for a toxic gain of function associated with various ALS-associated mutants of SOD1.

Beckman, Koppenol, and co-workers (38, 39) have also proposed that intact human SOD1 can activate PN to nitrosylate protein tyrosine residues. The mechanism would involve Cu(II)-catalyzed heterolytic cleavage of PN into the nitronium cation and CuO, with the former being a potent nitrosylating agent. Indeed, Lei et al. (379) demonstrated that there is a diminished or blocked protein nitration in Sod1−/− mice treated with APAP. They proposed that the block of hepatic protein nitration in those mice might partially explain their resistance to the APAP overdose. Adding functional holo-SOD1, but not apo-SOD1, to liver homogenates of the Sod1−/− mice mixed with a bolus of PN indeed resulted in increased protein nitration (758).

Likewise, GPX1 bears prooxidant potential. Several groups have demonstrated that this enzyme can aggravate nitrosative stress in mouse models (199, 343, 376, 377, 379, 458). This effect opposes the role of GPX1 in preventing nitrosative stress by catalyzing reduction of PN into nitrite using reducing equivalents from GSH (610). Although the precise mechanisms remain poorly understood (191, 197), attenuated protein nitration should help explain the increased resistance of Gpx1−/− hepatocytes to PN toxicity and lack of potentiation or even protection conferred by Gpx1 knockout against APAP hepatotoxicity (199, 343, 376, 377, 379, 458).

2. “Unwanted” modulation of reducing equivalents

Excessive enzymatic removal of ROS/RNS may lead to other detrimental downstream effects. For example, consumption of GSH as a reductant substrate depletes cells of GSH and thus outweighs the benefits of GSH-dependent ROS scavenging enzymes (199, 379, 458). Although uncatalyzed reduction of peroxide by GSH is slow (700), GPX1 is very efficient at catalyzing this reaction (235). However, GSH can directly scavenge other more reactive species, like hydroxyl radicals, HOCl, PN, and carbonate radicals (235). It can also regenerate antioxidants vitamins C and E. This may partially explains why overproduction of GPX1/Gpx1 can result in greater sensitivity to the destructive reactivity of APAP metabolites (199, 379, 458).

Elevating GSH may also be detrimental via mechanisms that involve S-glutathionylation and inactivation of various key proteins. For instance, GAPDH (462), eNOS (96, 431), certain tyrosine phosphatases (1), MAPK phosphatase 1 (331), mitochondrial thymidine kinase 2 (630), and protein disulfide isomerase (715) have all been reported to be inactivated by glutathionylation. This may either protect such enzymes from further damage, or can inhibit their function. While the precise pathways and mechanisms are yet unclear, NRF2 is emerging as a major regulator of oxidative and reductive extreme conditions in metabolism (57, 321). Upon a rise of ROS levels above normal, NRF2 helps to upregulate GSH synthase and GSSG reductase, G6PD of the pentose phosphate pathway, as well as antioxidant enzymes like TrxR1, SOD, and catalases. While initially activated to protect against oxidative stress, hyperactivity of NRF2 can however result in a shift towards reductive stress, due to overabundance of antioxidant factors and GSH that can lead to cardiomyopathy and hypertrophy (543). It is possible that the detrimental effects of reductive stress can be associated with distorted S-glutathionylation (227) and/or inappropriate suppression of critical ROS-dependent signaling pathways. The effects of antioxidant enzyme overexpression in this context remain to be better characterized.

3. Stress source and intensity-dependent roles

As ROS scavengers, both Gpx1 and Sod1 protect mice against the lethality and toxicity caused by ROS-generating diquat and paraquat (111, 113, 168, 199). However, the opposite is true when mice are treated with the RNS-generating APAP and kainic acids (309, 379). Indeed, the ultimate metabolic outcome from overexpression or knockout of a particular antioxidant enzyme should be decided by how the enzyme will alter the relative production and fate of ROS and RNS in a given context. Good examples are the impacts of SOD overproduction on cardiovascular diseases or myocardial ischemic injuries. First, elevated SOD may help to preserve NO bioavailability, by preventing its reaction with O2 to form PN, and thus allow NO to serve as a vessel relaxation factor to protect the cardiac function and survival (83, 493). On the other hand, the hyperactivity of SOD may promote formation of H2O2 which then triggers downstream signaling responses that may inhibit vascular pathogenesis (747). However, the role of vascular H2O2 can also depend on the location, as exemplified with H2O2 derived from overproduced SOD3 anchored to endothelial cells, which promotes VEGFR2 signaling and then potentially aggravates angiogenesis-dependent vascular diseases (508).

Knockouts of Txnrd1, Gpx1, and Sod1 produce different phenotypes of glucose and lipid metabolism in mice (302, 680, 684). While the knockout of Sod1 elevates endogenously derived intracellular O2, the mice display similar impairments in islet physiology, but distinct signaling mechanism compared with the Gpx1−/− mice with elevated intracellular peroxides. As shown in Figure 7, Sod1 knockout downregulates Pdx1 at three levels: epigenetic, gene, and protein, whereas the Gpx1 knockout affects only the Pdx1 protein. Interestingly, both knockouts suppress GSIS by elevating Ucp2 expression with decreased ATP production and affecting the mitochondrial potential in islets (684). Interestingly, only the GPX mimic ebselen, but not the SOD mimic copper diisopropylsalicylate (CuDIPs), rescues impaired GSIS in islets of all test genotypes including Gpx1−/− and Sod1−/− (684). The effects of ebselen seemed to be mediated via Pgc-1α while, in contrast, CuDIPs improve insulin secretion only in Sod1−/− islets with suppressed gene expression of the Pgc-1α pathway (685). These results demonstrate that Gpx1 and Sod1, via their respective ability to modulate different species of ROS, can differentially affect redox-sensitive pathways in regulating GSIS. However, the “sensor” in the target signal molecules or regulators that distinguish and react with the local changes of O2 and peroxide needs unveiling.

Even for the same oxidative insult, the necessity and mechanism of a given antioxidant enzyme will vary with stress intensity and antioxidant status. When mice are injected with a high dose of paraquat (50 mg/kg body wt) or diquat (24 mg/kg), Gpx1 becomes absolutely essential to promote mouse survival by protecting against depletion of NADPH and redox collapse (113, 195). In contrast, Gpx1−/− mice tolerate low doses of paraquat (12.5 mg/kg) and diquat (6 mg/kg) well if they are fed adequate selenium to saturate the expression of other selenoproteins (111, 196). Still, minute amounts of Gpx1 activity, raised by injection of selenium in selenium-deficient mice, become protective against a moderate dose of paraquat-induced hepatic necrosis and apoptosis (112). Comparatively, knockout of Sod1 in the mice enhances sensitivity to oxidative injury induced by a similar dose of paraquat (10 mg/kg) (264), implicating that generation of O2 contributes more to the paraquat-induced oxidative toxicity.

4. Compensatory inductions

“Hidden helpers” or compensatory responses induced by altering the expression of antioxidant enzymes may also help to explain the mechanisms of the observed phenotypes. The protection conferred by the Sod1 knockout against the APAP toxicity is associated with a 50% activity reduction in a key APAP-biotransforming enzyme, CYP2E1 (379), which catalyzes the formation of toxic metabolites of APAP. The Sod1 knockout also results in a 40% reduction of Gpx activity (684), which may also add to the resistance of the Sod1−/− mice to the drug overdose (756).

As elaborated above, NRF2 is a redox-sensitive transcription factor that controls protective responses to oxidative stress (419, 463). The protein is normally sequestered in the cytosol by Keap1 and marked for proteasomal degradation. Under oxidative stress, Keap1 becomes oxidized and NRF2 can then translocate to the nucleus where it initiates transcription of selected antioxidant enzyme genes (419, 496). In the Sod1−/− mice, a greater fraction of Nrf2 is localized in the nucleus and upregulates gene expression of many antioxidant proteins including glutathione S-transferases, sulfiredoxins, TrxR1, GSH synthases, and other reductases (239). The upregulation of these enzymes provides an increased antioxidant capacity in Sod1−/− mice against the APAP-derived oxidative stress. Likewise, the liver specific knockout of Txnrd1 enhances mouse resistance to the APAP toxicity by upregulation of the Nrf2-target genes and proteins, with more robust GSH biosynthesis, glutathionylation, and glucuronidation systems (302, 520). The increased resistance to acute lung injury induced by endotoxin in acatalasemic mice (759) results from the H2O2-mediated downregulation of cytokine expression in macrophages via inhibition of NFκB activation. With these compensatory mechanisms revealing intricate pathways of physiological coordination to cope with redox imbalances, caution should be given to evaluate functions of antioxidant enzymes as absolute or isolated entities, because they will always be context-dependent.

5. Overlapping and coordinated functions

CAT and GPX1 are two major antioxidant enzymes that are both responsible for removal of H2O2 although via distinct mechanisms (94, 184). Double-knockout mice deficient in both of these enzymes were generated to reveal insights into the overlap between these potentially redundant H2O2-detoxification systems (333). Interestingly, hepatic lipid peroxidation is not elevated in mice deficient in Gpx1 alone compared with that of wild-type mice (113), yet it is increased in mice lacking both Gpx1 and Cat (333).

In other cases, intrinsic expression of multiple isoforms of the same antioxidant enzymes in cells makes interpretations of oxidant-mediated diseases difficult. GPX1 is ubiquitous in all types of cells and GPX2 is in epithelium of the gastrointestinal tract (60, 117, 177). Gpx1−/− and Gpx2−/− mice are grossly normal. However, Gpx1−/−Gpx2−/− mice develop spontaneous ileocolitis and intestinal cancer (118, 174, 175). This occurrence of cancer is also associated with an increased rate of mutation in the intestine (372). Collectively, these results suggest that the two enzymes cooperatively attenuate intestinal flora-induced inflammation by removing H2O2 and alkyl hydroperoxides, thereby suppressing the vicious cycles of the inflammatory response and oxidant-mediated mutations and cancer.

Overproduction of either SOD2 or Cat, each having its distinct target of ROS, in pancreatic β cells of mice significantly delays but does not prevent the onset of diabetes induced by STZ compared with wild-type mice (99). However, the STZ-triggered increase in blood glucose is more effectively attenuated by overexpression of both enzymes in mice, suggesting that both O2 and H2O2 contribute to the dysfunction and death of pancreatic β cells caused by STZ. In contrast, double knockouts of Sod1 and Gpx1 did not produce more aggravated effects than the single knockout of Sod1 on mouse susceptibility to prooxidant toxicity, loss of islet β-cell mass and insulin synthesis, and dysregulation of glucose metabolism (376, 684, 758). Thus overlapping roles or coordinations between antioxidant enzymes are not a universal feature.

Double knockouts of antioxidant enzymes can also yield unexpected novel insights into mammalian redox control. This was recently exemplified when Txnrd1 was conditionally deleted from hepatocytes in mice lacking a functional Gsr gene, thus leading to livers lacking both of the two major cytosolic NADPH-dependent oxidoreductases TrxR1 and GR, presumed to be required for essentially all NADPH-dependent reductive activities in the cytosol. These mice were, surprisingly, found to be both viable and fertile. The reductive power was instead supplied solely by dietary methionine that became converted to GSH, which was likely used in single-turnover reactions and thus these livers avoid the reliance on NADPH (171).

6. Non-redox functions

SOD1 has been shown to play roles in copper and zinc metabolism (132, 690, 709). In Baker's yeast, overexpression or deletion of SOD1 affects the cell resistance or sensitivity to copper and zinc toxicity or deprivation, respectively (132, 709). More recent work suggests that SOD1 is also important for communicating the cellular stress response to low zinc (278, 709). Most interestingly, Tsang et al. (656) have demonstrated a role for SOD1 in cell signaling independent of its role in O2 disproportionation. They found that under oxidative stress, SOD1 translocated to the nucleus where it acted as a transcriptional activator of genes involved in oxidative resistance and repairing. Indeed, yeast cells expressing an allele of SOD1 that cannot get into the nucleus are more sensitive to oxidative stress. SOD1 represents ∼1% of total cellular protein (∼10–50 μM), and <1% of total SOD1 enzyme may be needed to protect against various oxidative insults (127, 557). Thus the recently discovered novel functions of SOD1, besides O2 scavenging, help explain its high cellular abundance and perhaps its paradoxical roles.

Another antioxidant enzyme that exhibits non-redox functions is PRX1 (307). The enzyme forms oligomeric species that exhibit chaperone activity upon oxidation of certain Cys to sulfinic acid. Such a mechanism enables the PRX family enzymes, which are better at scavenging low concentrations of peroxide, to be converted to chaperones to ensure proper protein folding under severe oxidative stress and high peroxide fluxes (409, 560, 705). Undoubtedly, the discovery of SOD1 as a novel transcriptional activator and PRX1 as a chaperone offers a new direction to elucidate the underlying mechanism for paradoxical roles of antioxidant enzymes. Additional non-redox functions of antioxidant enzymes include the role of Gpx4 as a structural protein in sperm mentioned above, or the cytokine-like properties of extracellular Trx1 or Trx80. It is possible, perhaps even likely, that additional non-redox related functions of classically considered antioxidant enzymes will be discovered, which should help in interpreting the phenotypes seen upon their genetic deletion or overexpression.

C. Metabolic Context and Reaction Environment Affecting Roles of Antioxidant Enzymes

1. Physiological versus pathophysiological conditions

Antioxidant enzymes may exert different impacts in physiological compared with pathological processes. While overexpressing Gpx1 induces type 2 diabetes-like phenotypes in mice without diabetic or obese-prone genetic background (445), the β cell-specific overexpression of GPX1 in ob/ob mice actually, in stark contrast, reverses hyperglycemia and improves β-cell volume and granulation (244). Similarly, knockout of Gpx1 impairs insulin synthesis and secretion in mice fed the normal diet (684), but enhances mouse resistance to a high-fat diet-induced insulin resistance (406). Therefore, roles of antioxidant enzymes under “normal” and “diseased” conditions should not be extrapolated or inferred from each other.

2. Temporal dependence of physiological effects

Short-term benefits of antioxidant enzyme alterations may lead to long-term harms, and vice versa. Indeed, overexpression of Gpx1 alone or in combination with SOD1 and SOD3 protects mouse islets from oxidative injury and improves islet graft function (480). However, the long-term overproduction of Gpx1 results in hyperinsulinemia, insulin resistance, and obesity (684). In contrast, knockout of Sod1 offers extra resistance to APAP overdose and insulin sensitivity in the young adult mice (379, 684), but leads to hepatocarcinogenesis in later life (167). When the hippocampal long-term potentiation (LTP), one of the major cellular mechanisms for learning and memory ability, is impaired in young (2 mo old) SOD1 overexpressing mice (201), the aged (2 yr old) transgenic mice actually exhibit an enlarged LTP (316) and consequently a better performance in spatial memory (315) compared with the wild-type mice. These differences implied a strong age dependence for the effects of the Sod1 deficiency and/or Sod1-derived peroxide based on the brain function.

3. Subcellular location-dependence effects

Subcellular localizations of particular antioxidant enzymes have profound effects on their roles, which need to be considered in interpretations of the mechanistic results. Overexpressing extracellular GPX3 protects mice from the APAP toxicity, while the overexpression of intracellular GPX1 sensitizes mice to the toxicity (458). Likewise, the β-cell specific overexpression of cytoplasmic Cat and the metallothionein gene, but not the mitochondrial Sod2, accelerates the cyclophosphamide-induced and spontaneously developed diabetes in the nonobese diabetic male mice (390). Overexpression of CAT in mitochondria, but not in the peroxisomes or nuclei, extends the median and maximal lifespan of the mice by 20% (584). This indicates that the interactions between ROS/RNS and their metabolizing enzymes should not be extrapolated from different subcellular compartments.

Likewise, peroxide derived from the yeast SOD1 protein that is proximal to a membrane-bound casein kinase is required to regulate energy metabolism in response to oxygen and glucose availability (557). The yeast SOD1 protein that is not targeted to the cytosol, or other SOD isoforms that are targeted to the cytosol like mitochondrial SOD2 or Candida albicans SOD3 are unable to regulate casein kinase signaling. Similarly, a small fraction of cytosolic PRX1 and PRX2 is associated with lipid rafts proximal to NADPH oxidase enzymes. Only the lipid raft -PRX1, but not cytosolic PRX1, is found to be phosphorylated at Tyr194 by a protein tyrosine kinase (PTK) of the Src family when cells are stimulated by growth factors (705). Phosphorylation of PRX1 near membranes has the effect of inactivating the enzyme, which promotes peroxide-mediated signals to propagate.

D. Cell Compartmentalization and Tissue Heterogeneity of Transgenes

1. Effects of cell types on responses toward similar changes of antioxidant enzymes

While the cardiac-specific overexpression of CAT/Cat generates many benefits for prolonging lifespan and protecting against cardiac injuries (208, 317, 660, 712, 748, 749), the same specific overexpression in the endothelium shows little protection against myocardial or vascular ischemia/reperfusion injury (704). In either tissue-specific overexpression of a given antioxidant enzyme, such as catalase in cardiomyocytes and pancreatic β cells (319, 717), or ubiquitous overexpression of an antioxidant enzyme in mice, the intended overexpression may be very heterogeneous in different types of cells within an organ. Likewise, extents of a global overexpression of a transgene in mouse tissues may be restricted to certain organs, but not as widely spread as the corresponding endogenous mouse genes or the genes whose promoters are used in the transgene constructs (such as the human β-actin promoter) (266, 494, 509, 716, 727).

The heterogeneity of transgene expression cannot be appropriately assessed when homogenate of the entire organ is used for expression study. To circumvent this problem and utilize endogenous “native” gene regulation, large genomic fragments containing the genes of interest have been used to overexpress SOD1 and CAT (103). However, whether the specificity of transgene expression can be applied to each individual type of cells within each organ is still an open question.

Heterogeneity of transgene expression is also shown even in different mice carrying an identical transgene. For example, the same 14.5-kb genomic fragment containing the entire human SOD1 gene has been used independently by several laboratories to generate transgenic mice (87, 170, 231, 681). Although the SOD1 overexpression protects the heart against an in vitro model of ischemia/reperfusion in two independent lines of transgenic mice, the cell specificity of overexpression in these mice is quite different. The gene is overexpressed in both endothelial cells and cardiomyocytes in one line of transgenic mice (681), but exclusively in coronary vascular cells including endothelial cells and smooth muscle cells but not cardiomyocytes in another line (106). Therefore, immune-histochemical studies are needed to identify the types of cells expressing the transgene in the target organs, and more than one line of transgenic mice carrying the same transgene should be employed in physiological studies to ensure reproducibility of the experiments. The latter approach is even more critical when homozygous transgenic mice are used in the experiments, because the transgene occasionally disrupts the expression of a normal mouse gene at the site of integration by the mechanism referred to as “insertional mutagenesis,” leading to a phenotype that is unrelated to the expression of the transgene (708). As a given antioxidant enzyme may not be sufficiently overexpressed in targeted cells within an organ that are vulnerable to a particular oxidant-mediated injury, a negative result does not rule out the enzyme function in defense against the injury in those cells.

The tissue heterogeneity of transgene expression may also affect human implications of findings from a particular animal model. Noteworthy, SOD3 in human aorta accounts for ∼50% of the total SOD activity, whereas the enzyme in rat aorta represents only 5% of the total SOD activity due to a key amino acid difference that affects tissue binding in vessels (178, 619). As a result, the rat essentially lacks vascular SOD3 and, consequently, the observed protection of SOD3 against vascular diseases in rat models may be easily over-interpreted.

2. Genetic background of mouse models

Most transgenic and knockout mice are initially generated in a mixed genetic background (272, 515), and it will take 10–12 generations of backcrossing to become congenic. Because this crossing may take several years, most of the phenotypic studies, at least initially, are performed on mice in a mixed genetic background. Such studies should be interpreted with caution, since the genetic background of the mice may contribute to the observed phenotypes. For example, strain C57BL/6J (B6) mice are more susceptible to hyperoxia-induced lung injury than C3H/HeJ (C3) mice (287). Further studies using linkage analysis have shown that the B6 mice carry a nucleotide substitution in the promoter region of the Nrf2 gene (116). This Nrf2 polymorphism cosegregates with the susceptible phenotype of B6 mice to hyperoxia. Therefore, when SOD2 transgenic mice in a B6 × C3 mixed genetic background are used for study of hyperoxia-induced lung injury, tolerance to exposure is determined by both the origin of the Nrf2 allele and expression of the SOD2 transgene (266). Therefore, control experiments should be conducted for functional studies in mice with the identical genetic background, preferably littermates of the experimental mice.

3. Heterozygous mouse models and human relevance

To date, most of the phenotypic studies have been performed using homozygous knockout mice (if viable) compared with wild-type mice. However, studies using heterozygous mice with a partial deficiency may be more relevant to human diseases, since humans being fully devoid of a protein or enzyme are likely to be rare. Although relatively limited studies have documented the phenotypes of such mice that express approximately one-half of the normal amount of enzyme, some results are intriguing. For example, under normal physiological conditions, the time to development of malignant tumors in Prx1+/− mice is between those of Prx1−/− and wild-type mice. In addition, hemolytic anemia was first observed in the Prx1+/− mice at 12 mo of age compared with 9 mo of the null mice, whereas wild-type mice are free of this disease (484). Thus a partial deficiency in Prx1 results in phenotypes being intermediate between complete deficiency and normal in mice. On the contrary, while Sod1−/− females show a declined fertility (489), the fertility of the Sod1+/− females are normal (264, 443). In response to trauma-induced dysfunction of mitochondrial respiration in brain, Cat+/− mice are as vulnerable as Cat−/− mice (268). In contrast, the phenotype of Sod1+/− mice resembles that of wild-type mice in response to acute paraquat toxicity (10 mg/kg body wt) (264). Therefore, the effect of a partial deficiency in antioxidant enzyme on untreated mice and oxidant-mediated disease models varies from gene to gene. While future research on the physiological role of antioxidant enzymes should consider more partial knockdown or knockout models, current implications from the homozygous knockout mouse models need to be verified in human studies.

In summary, we have postulated a series of mechanisms in this section that should underpin the “paradoxical” outcomes of antioxidant enzyme deletion or overexpression. Figure 13 highlights the central concept that effects of antioxidant enzyme modulation arise from a complex interplay between the activities of the antioxidant enzymes with their ROS/RNS substrates (and reductants such as GSH or NADPH), as well as the importance of the environmental context in which they operate. It is our hope that this figure, along with our deliberations, prompt readers to recognize that the mechanisms by which nature masterfully orchestrates these seemingly paradoxical events are evolved to maintain redox homeostasis and are critical towards understanding both health and disease.

FIGURE 13.

FIGURE 13.Scheme of mechanisms of paradoxical outcomes upon modulation of antioxidant enzyme status. In general, either detrimental or beneficial impacts of antioxidant enzyme overexpression or knockout arise from a complex interplay among redox active enzymes, their substrates, and the enzymatic reaction environment. Different ROS/RNS species and antioxidant enzymes discussed in this review are illustrated, along with their representative features of chemistry, free radical biology, and metabolism that may all trigger paradoxical outcomes. Specifically, the dose, reactivity, and localization of ROS/RNS substrates can lead to differential impacts on oxidative stress and redox signaling pathways. Impacts and mechanisms of reductant substrates (e.g., GSH) in the “paradox” are shown in the context of antioxidant enzyme catalysis. The antioxidant enzymes can themselves contribute to the paradoxical outcomes by acting as pro-oxidants, either by catalyzing production of certain ROS/RNS or overconsuming reducing equivalents, depleting ROS/RNS required for signaling, acting on noncanonical substrates, exhibiting nonredox functions, inducing compensatory responses, or having overlapping functions with other enzymes. The environmental context, i.e., physiological (nonstress) or pathophysiological (metabolic stress, oxidative injury, nutrient deficiency, or drug toxicity) state, the experimental model, as well as spatial or time constraints, will determine the final phenotype. Thus apparent paradoxes in antioxidant enzyme overexpression and knockout studies should be viewed in a well-defined physiological context as combined interactions of all of these factors. Cat, catalase; Gpx, glutathione peroxidase; Grx, glutaredoxin; MsR, methionine sulfoxide reductase; Prx, peroxiredoxin; ROS, reactive oxygen species; RNS, reactive nitrogen species; Sepp1, selenoprotein P; Trsp, selenocysteine tRNA gene; Sod, superoxide dismutase; Trx, thioredoxin; TrxR, thioredoxin reductase.


V. HEALTH AND NUTRITION IMPLICATIONS

A. Antioxidant Enzymes in Relation to Human Diseases

Catalase-deficient patients, classified as acatalasemic or hypocatalasemic, are found in many countries (495). These patients can have different alterations of the catalase gene including substitution (692, 693), deletion (262), and insertion (222, 225). Being apparently healthy, patients with acatalasemia may display increased risks of a progressive oral gangrene (166, 637, 638) and type 2 diabetes mellitus, especially in females (223). Still, the rather common occurrence of this autosomal recessive disease and its mild symptomatology suggests that catalase has mainly redundant activities with regards to human H2O2 removal pathways.

Two well-known neurodegenerative diseases: familial ALS and Down's syndrome, exemplify the significant health implications of antioxidant enzymes in a “paradoxical” manner. While dominantly inherited mutations of SOD1 gene account for 20% of the familial ALS cases (569), the pathophysiology seems to be due to a gain of mutant protein toxicity independent of the normal enzymatic activity of SOD1. Several lines of transgenic mice expressing Sod1 mutants have indeed displayed pathological characteristics reminiscent of those seen in ALS (67). The Down's syndrome patients usually display a 50% increase in SOD activity (14, 131, 612). Although transgenic mouse models have been developed for this disease (23, 24, 524, 580), the underlying mechanisms of SOD1 toxicity in Down's syndrome are not understood (143, 366). In addition, mutations of SOD2 in humans are associated with idiopathic cardiomyopathy, sporadic motor neuron defect, and cancer (261). However, Sod2−/− mice generated by targeted disruption only partially recapitulate these human symptoms; rather, these mice display metabolic phenotypes including fatty liver and cardiomyopathy (388). Indeed, polymorphisms of SOD enzymes, catalase, and GPX1 have been implicated in association with a number of human metabolic disorders such as diabetes and cardiovascular diseases, as well as cancers (reviewed in Refs. 130, 253).

Altered nutritional selenium intake has long been implied in several diseases that are believed to be explained mainly by aberrant selenoprotein functions (554). The first examples of genetically and molecularly defined diseases of insufficient selenoprotein synthesis were found to relate to mutations in the selenocysteine insertion sequence binding protein-2 (SBP2) involved in translational insertion of Sec into selenoprotein and leading to complex diseases with hypothyroidism as a main symptom (25, 150, 582). These patients are however only partially deficient in selenoproteins and considering that deletion of Trsp and some of the selenoproteins in mice is embryonically lethal (see above), it is unlikely that patients would survive with a total lack of selenoprotein synthesis, but additional polymorphisms and other aberrations in specific selenoproteins are likely to be discovered in relation to disease.

Among the genetic variants of GPX enzymes, GPX1Pro198Leu polymorphism is the most studied case. In a small randomized trial with 37 morbidly obese women (BMI >45), this variant precluded the protection against DNA breaks by daily supplementation of one Brazil nut daily (290 μg selenium/day) for 8 wk (121). Furthermore, the same variation is associated with decreased selenium status in Alzheimer's patients (75), lowered GPX activity and increased breast cancer risk in Danish women (553), predisposition to colorectal adenomas or carcinomas based on the Norwegian cohort NORCCAP (241), and increased prevalence of cardiovascular disease on the cohort of 184 Japanese with type 2 diabetes (238). These associations appear to be specific, as no such relationship was found between the same variant and the risk of basal cell carcinoma in the cohort of 317 Danish (677). Another GPX1 variant (C198T) lowering the enzyme activity was identified in the South Indian population, which resulted in increased incidences of type 2 diabetes (C/T, 1.4-fold; T/T, 1.8-fold) (547).

Several single nucleotide polymorphisms on GPX2 are found to affect Barrett's esophagus and esophageal adenocarcinoma (479). Polymorphisms of GPX3 are known to suppress the expression of this gene and serve as a risk factor for thrombosis in cerebral veins (676). In the 3'-untranslated region of GPX4, the T/C variation at position 718 is linked to cancer susceptibility, with the T variant being associated with a lower risk for developing colorectal cancer (44). Likewise, polymorphisms in transcription factor binding sites of the PRX6 promoter are associated with less favorable overall survival in breast cancer patients (589).

B. Novel Treatments of Antioxidant Enzyme-Related Diseases

The impact of antioxidant enzymes in disease may possibly offer novel treatment options for redox-related diseases, provided that the molecular mechanisms are known and can be specifically targeted. RNA interference (RNAi) technologies may thus possibly be developed for treatment of ALS originated from single nucleotide polymorphisms in SOD1 (472, 569). Mice expressing the missense mutant SOD1G93A-targeting shRNA were created to prove the principle of therapeutic potential (154, 544, 552, 566, 713).

Commonly used drugs for treating cardiovascular diseases, such as β-adrenocepter blocker carvedilo, angiotensin converting enzymes, and statins (2, 738), bear SOD-like activities that suppress O2. A GPX mimic may be used to improve GSIS impaired by the GPX1 deficiency (685). Overexpressing one or several antioxidant enzyme genes proves effective to prolong the survival of islet graft against the anticipated host oxidative attack (480). When large doses of chemotherapeutic agents or radiation induce severe oxidative stress, treating the patients with antioxidant enzyme mimics may help restore their redox homeostasis (359).

Meanwhile, there have been many studies aiming for virally mediated approaches to increase expression of antioxidant enzymes for protective effects in models of hypertension, restenosis, myocardial infarction, stroke, and other diseases (see Ref. 711 for a review). For example, a gene delivery of antioxidant enzymes such as GPX1 and SOD1 was shown to attenuate oxidative stress in the brain of rodent models of HIV-associated neurocognitive disorders, Parkinson's disease, and diabetic complications (5, 6, 453, 564). Similarly, gene delivery for expression of SOD3 protects against the monocrotaline-induced hypertension in the lung of rats (312). However, the safety and efficacy of gene therapy are still a concern, and therapeutic potentials of viral delivery of antioxidant enzyme genes to specific tissues remain an open question.

In contrast, inhibiting a given antioxidant enzyme or specifically silencing its gene expression may help treat disorders related to a gain of enzymatic function. As stated above, there is a great potential of using RNAi to specifically suppress the toxic mutant of Sod1 gene associated with ALS (154, 544, 552, 566, 713). In addition, microRNA (miRNA) species, regulators of mRNA stability and translation, have been recently proposed as biomarkers for a variety of diseases (449). Although numerous miRNAs targeting antioxidant enzymes have been identified in cultured cells (243, 682), little is known about the reciprocal interactions between antioxidant enzymes and miRNA expression during pathogenesis.

Many types of drug-resistant cancer cells express high levels of antioxidant enzymes such as SOD, GPX, TrxR1 and PRX (17, 64, 517). Pretreating these cells with specific antioxidant enzyme antagonists or genetically silencing the target gene shall improve the anti-cancer drug efficacy (666, 741). Similarly, preconditioning the antioxidant enzyme status may help minimize toxicities of some commonly used drugs. Theoretically, hepatotoxicity of APAP may be attenuated if the patients are treated with TrxR1, SOD1, or GPX1 inhibitors, perhaps along with some GPX3 mimic, before administration. This notion is based on the fact that knockout of Txnrd1, Sod1, or Gpx1 renders mice resistant to the drug-induced protein nitration and toxicity (see above), but overexpression of GPX3 protects against APAP hepatotoxicity (458).

Targeting of antioxidant enzymes may possibly also be applied to treat chronic diseases such as type 2 diabetes. Insulin resistance is the hallmark of the disease and is inversely related to the oxidative inhibition of protein phosphatases in GSIS. When Gpx1 overexpression diminishes intracellular H2O2 and lifts the oxidative inhibition of protein phosphatases, causing insulin resistance (445), knockout of Gpx1 and Sod1 alone or together improved insulin sensitivity via the opposite mechanism (684). Therefore, the injected insulin could be more effective in lowering blood glucose if GPX1 and SOD1 are temporarily downregulated prior to insulin administration. Clearly, such clinical protocols based on findings from mouse experiments need to be studied and duly verified in human studies.

C. Antioxidant Nutrients

1. Perception and mixed outcomes of intervention trials

Antioxidant nutrients in foods often refer to vitamins C and E, carotenoids (particularly β-carotene), and certain trace elements such as selenium and zinc. Antioxidants are widely used to preserve food and beverages and to promote value-added product sales because of their perceived health benefits (182, 233). Indeed, many people believe that “antioxidant is good, more antioxidant is better” (234). However, more than 100 nutritional intervention trials conducted during the past 20 years (4547) have shown disappointing outcomes of administering high or pharmacological doses of dietary antioxidant nutrients (45, 46, 221, 233, 235, 537). In contrast, supra-nutritional intake of selenium or elevated serum selenium concentrations seem associated with increased risk of type 2 diabetes (9, 119, 134, 363, 620). Although a reanalysis of the data from the large Selenium and Vitamin E Cancer Prevention (SELECT) trial (400) found the risk for increased prevalence of diabetes to be attributed to vitamin E supplementation (340), this controversial finding underscores the potential risk of overdosing antioxidant nutrients. It also points out the need for a thorough understanding of selenium biology before large nutritional selenium trials are initiated or when their results are to be interpreted (249, 555).

2. Mode of actions by antioxidant nutrients

Antioxidant nutrients may contribute to overall antioxidant defense and interact with antioxidant enzymes in several ways. First, some of these nutrients like vitamins C and E directly scavenge ROS and RNS. Second, some of them serve as cofactors of antioxidant enzymes. Examples include selenium in the form of Sec in selenoproteins, copper, zinc, and manganese in SOD as well as iron in catalase. Dietary selenium deficiency is related to several diseases as is selenium toxicity (554). While iron and zinc deficiencies are quite common, deficiencies of manganese and copper are rare in humans. However, supplementing these nutrients to adequate subjects does not likely elevate their pertaining antioxidant enzyme activities because the activities are supposed to be saturated by those nutrients at the nutrient requirement levels. That fact may partially explain the lack of positive effects of long-term supplementation of antioxidant nutrients in adequate subjects. Third, antioxidant nutrients regulate antioxidant enzyme gene expression and protein production. For example, dietary vitamin E seems to downregulate certain selenoprotein gene expressions (282) and upregulate SOD activity (506). Optimal intakes of antioxidant nutrients for the balance between body antioxidant enzymes and ROS/RNS still remain elusive.

3. Effects of phytochemicals on antioxidant enzymes

Plant foods contain a diverse range of secondary metabolites of bioactive molecules (phytochemicals) (380). Although these low-molecular-weight compounds do not seem to decrease systemic oxidative damage, polyphenols, carotenoids, and tocopherols may reach high concentrations in the gastrointestinal tract and exert effects there (236, 237). Moreover, some of these phytochemicals (e.g., polyphenols) can exert prooxidant effects.

As discussed above, NRF2 plays a key role in maintenance of cellular redox homeostasis under oxidative stress (49, 328, 464). Phytochemicals such as EGCG, curcumin, and isothiocyanates may induce oxidative or covalent modification of thiols in cysteine residues of regulators such as Keap1, resulting in dissociation of NRF2 from Keap1 and its translocation to the nucleus where NRF2 can regulate gene expression of more than 200 antioxidant and phase II detoxifying enzymes (74, 252, 645). Thus using naturally occurring phytochemicals to upregulate NRF2 may be a strategy for preventing or treating chronic diseases due to insufficient NRF2 activities (49, 156, 214). However, many of these compounds also inhibit TrxR1 (89), and the resulting long-term impact on disease must be better understood before guidelines on prevention through supplementation with phytochemicals should be given.

NRF2 can exert different roles in effects of various phytochemicals on cancer prevention and development (29, 35, 145, 420, 460, 464, 488, 596). Many phytochemicals characterized as NRF2 inducers (751) can be either chemopreventive or oncogenic (618). This has promoted scientists to search for NRF2 inhibitors. For example, brusatol isolated from the seeds of Brucea sumatrana may inhibit NRF2 and enhance the efficacy of chemotherapeutic drugs in a mouse xennograft model (558). A coffee alkaloid trigonelline inhibits NRF2 and renders pancreatic cancer cells susceptible to anti-cancer drug-induced apoptosis (16). Meanwhile, NRF2 can act as a protooncogene (596), suggesting that protective effects of its activities might exist only in normal noncancerous cells and tissues. Overexpression of Nrf2 indeed causes chemoresistance (742), whereas NRF2 inhibitors can overcome it (488). The complex nature between interactions of phytochemicals with NRF2 will require a genuinely personalized use of such compounds for cancer prevention or treatment (49, 210).

VI. CLOSING REMARKS

Strict control of ROS and RNS at physiological levels is essential to avoid disease, with neither too much nor too little being good. Recently, James Watson (689) hypothesized that several chronic diseases such as diabetes, dementias, cardiovascular disease, and certain types of cancers may all be linked to a failure to generate sufficient ROS. Another, complementary, theory is the Triage theory proposed by Bruce Ames underscoring that distortions in trace element usage by age underpins several diseases, which thereby also includes effects on several antioxidant enzymes (10, 444).

In this review, we have attempted to provide comprehensive analyses of the paradoxical functions of SOD, catalase, GPX, TrxR, Trx, Grx, and PRX enzymes, along with other selenoproteins and selenoprotein synthesis-related Trsp, in metabolism, health, and disease. While “paradoxes” associated with these enzymes signify an alternative requirement for the body to maintain metabolic homeostasis, our understanding of the underlying mechanisms is far from clear. We do not know how antioxidant enzymes respond to demands for a tight control of their substrates or products in specific tissues or whole body. We know very little of functions of antioxidant enzymes independent of redox modulation. Their prooxidant catalytic potential and mechanism are not fully recognized or understood. Likewise, little is yet revealed regarding feedback mechanism of individual antioxidant enzymes and global coordination of different enzyme families in coping with various ROS/RNS-initiated events.

Because most of our discussions are based on animal experiments, many of the findings need to be verified in humans. It is clear that a number of human diseases are associated with genetic defects or polymorphism in specific antioxidant enzymes. However, specific, sensitive, and reliable indicators of in vivo redox status are yet explored to identify the optimal range of the antioxidant enzyme activities and ROS/RNS tone required by individuals according to personal genetic makeup, life-style, and living environment. While mechanisms outlined in this review for the paradoxical roles of antioxidant enzymes may lead to alternative therapy strategies, the challenge will be to identify surfeits and deficits among the complex array of given diseases to design the most effective treatment. Antioxidant nutrients and phytochemicals can affect production of ROS/RNS, functions of antioxidant enzymes, and the balance between the two. It remains to be determined when and how these supplements are beneficial, wasteful, or even detrimental. In conclusion, antioxidant enzymes and their ROS/RNS substrates represent pairs of natural complements. Basic mechanisms and clinical implications for their interdependence and counterbalance in physiology and health warrant intensive research.

GRANTS

Research related to this review was supported in part by National Institutes of Health (NIH) Grant DK 53018 and Natural Science Foundation of China (NSFC) Grant 31320103920 (to X. G. Lei); by NSFC Grants 31201065 and 31310103026 and Zhejiang Provincial Natural Science Fund for Distinguished Young Scholars ( LR13H020002 ) (to J. H. Zhu); by an award from the Cancer Prevention Research Trust , UK (to Y. P. Bao); by start-up support from The Georgia Institute of Technology and National Institute of Environmental Health Sciences of the NIH under award R21ES025661 (to A. R. Reddi); and by support from The Swedish Research Council , The Swedish Cancer Society , and Karolinska Institutet (to A. Holmgren and E. S. J. Arnér).

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the authors.

FOOTNOTES

  • Genes and proteins are in this article typically named according to HUGO (http://www.genenames.org) and MGI (http://www.informatics.jax.org/mgihome/nomen/gene.shtml) guidelines for human and mouse, respectively, unless other names or abbreviations are by convention more prevalent in the literature. In some cases mice have been studied with overexpression from human transgenes, which may be somewhat confusing in terms of nomenclature. We believe, however, that the references given in the tables of this review will serve as useful reference material for any reader interested in the exact gene constructs that are being discussed.

acknowledgments

Address for reprint requests and other correspondence: X. G. Lei, Dept. of Animal Science, Cornell University, Ithaca, NY 14853 (e-mail: ).

REFERENCES

  • 1. Abdelsaid MA, El-Remessy AB. S-glutathionylation of LMW-PTP regulates VEGF-mediated FAK activation and endothelial cell migration. J Cell Sci 125: 4751–4760, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 2. Adam O, Laufs U. Antioxidative effects of statins. Arch Toxicol 82: 885–892, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 3. Adluri RS, Thirunavukkarasu M, Zhan L, Akita Y, Samuel SM, Otani H, Ho YS, Maulik G, Maulik N. Thioredoxin 1 enhances neovascularization and reduces ventricular remodeling during chronic myocardial infarction: a study using thioredoxin 1 transgenic mice. J Mol Cell Cardiol 50: 239–247, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 4. Aesif SW, Anathy V, Kuipers I, Guala AS, Reiss JN, Ho YS, Janssen-Heininger YM. Ablation of glutaredoxin-1 attenuates lipopolysaccharide-induced lung inflammation and alveolar macrophage activation. Am J Respir Cell Mol Biol 44: 491–499, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 5. Agrawal L, Louboutin JP, Reyes BA, Van Bockstaele EJ, Strayer DS. Antioxidant enzyme gene delivery to protect from HIV-1 gp120-induced neuronal apoptosis. Gene Ther 13: 1645–1656, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 6. Agrawal L, Louboutin JP, Reyes BA, Van Bockstaele EJ, Strayer DS. HIV-1 Tat neurotoxicity: a model of acute and chronic exposure, and neuroprotection by gene delivery of antioxidant enzymes. Neurobiol Dis 45: 657–670, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 7. Ahmed MN, Codipilly C, Hogg N, Auten RL. The protective effect of overexpression of extracellular superoxide dismutase on nitric oxide bioavailability in the lung after exposure to hyperoxia stress. Exp Lung Res 37: 10–17, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 8. Ahmed MN, Zhang Y, Codipilly C, Zaghloul N, Patel D, Wolin M, Miller EJ. Extracellular superoxide dismutase overexpression can reverse the course of hypoxia-induced pulmonary hypertension. Mol Med 18: 38–46, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 9. Akbaraly TN, Arnaud J, Rayman MP, Hininger-Favier I, Roussel AM, Berr C, Fontbonne A. Plasma selenium and risk of dysglycemia in an elderly French population: results from the prospective Epidemiology of Vascular Ageing Study. Nutr Metab 7: 21, 2010.
    Crossref | Google Scholar
  • 10. Ames BN. Low micronutrient intake may accelerate the degenerative diseases of aging through allocation of scarce micronutrients by triage. Proc Natl Acad Sci USA 103: 17589–17594, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 11. Andreassen OA, Ferrante RJ, Dedeoglu A, Albers DW, Klivenyi P, Carlson EJ, Epstein CJ, Beal MF. Mice with a partial deficiency of manganese superoxide dismutase show increased vulnerability to the mitochondrial toxins malonate, 3-nitropropionic acid, and MPTP. Exp Neurol 167: 189–195, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 12. Andreassen OA, Ferrante RJ, Klivenyi P, Klein AM, Shinobu LA, Epstein CJ, Beal MF. Partial deficiency of manganese superoxide dismutase exacerbates a transgenic mouse model of amyotrophic lateral sclerosis. Ann Neurol 47: 447–455, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 13. Anestal K, Prast-Nielsen S, Cenas N, Arner ES. Cell death by SecTRAPs: thioredoxin reductase as a prooxidant killer of cells. PLoS One 3: e1846, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 14. Arbuzova S, Hutchin T, Cuckle H. Mitochondrial dysfunction and Down's syndrome. Bioessays 24: 681–684, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 15. Ardanaz N, Yang XP, Cifuentes ME, Haurani MJ, Jackson KW, Liao TD, Carretero OA, Pagano PJ. Lack of glutathione peroxidase 1 accelerates cardiac-specific hypertrophy and dysfunction in angiotensin II hypertension. Hypertension 55: 116–123, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 16. Arlt A, Sebens S, Krebs S, Geismann C, Grossmann M, Kruse ML, Schreiber S, Schafer H. Inhibition of the Nrf2 transcription factor by the alkaloid trigonelline renders pancreatic cancer cells more susceptible to apoptosis through decreased proteasomal gene expression and proteasome activity. Oncogene 32: 4825–4835, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 17. Arner ES. Focus on mammalian thioredoxin reductases—important selenoproteins with versatile functions. Biochim Biophys Acta 1790: 495–526, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 18. Arner ES, Holmgren A. Physiological functions of thioredoxin and thioredoxin reductase. Eur J Biochem 267: 6102–6109, 2000.
    Crossref | PubMed | Google Scholar
  • 19. Arner ES, Holmgren A. The thioredoxin system in cancer. Semin Cancer Biol 16: 420–426, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 20. Aslund F, Zheng M, Beckwith J, Storz G. Regulation of the OxyR transcription factor by hydrogen peroxide and the cellular thiol-disulfide status. Proc Natl Acad Sci USA 96: 6161–6165, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 21. Atkinson HJ, Babbitt PC. An atlas of the thioredoxin fold class reveals the complexity of function-enabling adaptations. PLoS Comput Biol 5: e1000541, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 22. Avissar N, Ornt DB, Yagil Y, Horowitz S, Watkins RH, Kerl EA, Takahashi K, Palmer IS, Cohen HJ. Human kidney proximal tubules are the main source of plasma glutathione peroxidase. Am J Physiol Cell Physiol 266: C367–C375, 1994.
    Link | ISI | Google Scholar
  • 23. Avraham KB, Schickler M, Sapoznikov D, Yarom R, Groner Y. Down's syndrome: abnormal neuromuscular junction in tongue of transgenic mice with elevated levels of human Cu/Zn-superoxide dismutase. Cell 54: 823–829, 1988.
    Crossref | PubMed | ISI | Google Scholar
  • 24. Avraham KB, Sugarman H, Rotshenker S, Groner Y. Down's syndrome: morphological remodelling and increased complexity in the neuromuscular junction of transgenic CuZn-superoxide dismutase mice. J Neurocytol 20: 208–215, 1991.
    Crossref | PubMed | Google Scholar
  • 25. Azevedo MF, Barra GB, Naves LA, Ribeiro Velasco LF, Godoy Garcia Castro P, de Castro LC, Amato AA, Miniard A, Driscoll D, Schomburg L, de Assis Rocha Neves F. Selenoprotein-related disease in a young girl caused by nonsense mutations in the SBP2 gene. J Clin Endocrinol Metab 95: 4066–4071, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 26. Bai J, Cederbaum AI. Overexpression of catalase in the mitochondrial or cytosolic compartment increases sensitivity of HepG2 cells to tumor necrosis factor-alpha-induced apoptosis. J Biol Chem 275: 19241–19249, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 27. Baliga MS, Diwadkar-Navsariwala V, Koh T, Fayad R, Fantuzzi G, Diamond AM. Selenoprotein deficiency enhances radiation-induced micronuclei formation. Mol Nutr Food Res 52: 1300–1304, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 28. Banning A, Deubel S, Kluth D, Zhou Z, Brigelius-Flohe R. The GI-GPx gene is a target for Nrf2. Mol Cell Biol 25: 4914–4923, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 29. Bao Y, Wang W, Zhou Z, Sun C. Benefits and risks of the hormetic effects of dietary isothiocyanates on cancer prevention. PLoS One 9: e114764, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 30. Bar-Peled O, Korkotian E, Segal M, Groner Y. Constitutive overexpression of Cu/Zn superoxide dismutase exacerbates kainic acid-induced apoptosis of transgenic-Cu/Zn superoxide dismutase neurons. Proc Natl Acad Sci USA 93: 8530–8535, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 31. Baran H, Loscher W, Mevissen M. The glycine/NMDA receptor partial agonist d-cycloserine blocks kainate-induced seizures in rats. Comparison with MK-801 and diazepam. Brain Res 652: 195–200, 1994.
    Crossref | PubMed | ISI | Google Scholar
  • 32. Barrett CW, Ning W, Chen X, Smith JJ, Washington MK, Hill KE, Coburn LA, Peek RM, Chaturvedi R, Wilson KT, Burk RF, Williams CS. Tumor suppressor function of the plasma glutathione peroxidase gpx3 in colitis-associated carcinoma. Cancer Res 73: 1245–1255, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 33. Barrett DM, Black SM, Todor H, Schmidt-Ullrich RK, Dawson KS, Mikkelsen RB. Inhibition of protein-tyrosine phosphatases by mild oxidative stresses is dependent on S-nitrosylation. J Biol Chem 280: 14453–14461, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 34. Barrett WC, DeGnore JP, Keng YF, Zhang ZY, Yim MB, Chock PB. Roles of superoxide radical anion in signal transduction mediated by reversible regulation of protein-tyrosine phosphatase 1B. J Biol Chem 274: 34543–34546, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 35. Bauer AK, Cho HY, Miller-Degraff L, Walker C, Helms K, Fostel J, Yamamoto M, Kleeberger SR. Targeted deletion of Nrf2 reduces urethane-induced lung tumor development in mice. PLoS One 6: e26590, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 36. Baumbach GL, Didion SP, Faraci FM. Hypertrophy of cerebral arterioles in mice deficient in expression of the gene for CuZn superoxide dismutase. Stroke 37: 1850–1855, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 37. Beck MA, Esworthy RS, Ho YS, Chu FF. Glutathione peroxidase protects mice from viral-induced myocarditis. FASEB J 12: 1143–1149, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 38. Beckman JS, Carson MC, Smith CD, Koppenol WH. ALS, SOD and peroxynitrite. Nature 364: 584, 1993.
    Crossref | PubMed | ISI | Google Scholar
  • 39. Beckman JS, Koppenol WH. Nitric oxide, superoxide, peroxynitrite: the good, the bad, ugly. Am J Physiol Cell Physiol 271: C1424–C1437, 1996.
    Link | ISI | Google Scholar
  • 40. Ben-Ari Y. Limbic seizure and brain damage produced by kainic acid: mechanisms and relevance to human temporal lobe epilepsy. Neuroscience 14: 375–403, 1985.
    Crossref | PubMed | ISI | Google Scholar
  • 41. Beni SM, Tsenter J, Alexandrovich AG, Galron-Krool N, Barzilai A, Kohen R, Grigoriadis N, Simeonidou C, Shohami E. CuZn-SOD deficiency, rather than overexpression, is associated with enhanced recovery and attenuated activation of NF-kappaB after brain trauma in mice. J Cereb Blood Flow Metab 26: 478–490, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 42. Bensadoun JC, Mirochnitchenko O, Inouye M, Aebischer P, Zurn AD. Attenuation of 6-OHDA-induced neurotoxicity in glutathione peroxidase transgenic mice. Eur J Neurosci 10: 3231–3236, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 43. Berg M, Bruhn T, Johansen FF, Diemer NH. Kainic acid-induced seizures and brain damage in the rat: different effects of NMDA- and AMPA receptor antagonists. Pharmacol Toxicol 73: 262–268, 1993.
    Crossref | PubMed | Google Scholar
  • 44. Bermano G, Pagmantidis V, Holloway N, Kadri S, Mowat NAG, Shiel RS, Arthur JR, Mathers JC, Daly AK, Broom J, Hesketh JE. Evidence that a polymorphism within the 3'UTR of glutathione peroxidase 4 is functional and is associated with susceptibility to colorectal cancer. Genes Nutr 2: 225–232, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 45. Bjelakovic G, Nikolova D, Gluud LL, Simonetti RG, Gluud C. Antioxidant supplements for prevention of mortality in healthy participants and patients with various diseases. Cochrane Database Syst Rev CD007176, 2008.
    Crossref | ISI | Google Scholar
  • 46. Bjelakovic G, Nikolova D, Gluud LL, Simonetti RG, Gluud C. Antioxidant supplements for preventjion of mortality in healthy participants and patients with various diseases. Cochrane Database Syst Rev CD007176, 2012.
    Crossref | ISI | Google Scholar
  • 47. Bjelakovic G, Nikolova D, Gluud LL, Simonetti RG, Gluud C. Mortality in randomized trials of antioxidant supplements for primary and secondary prevention: systematic review and meta-analysis. JAMA 297: 842–857, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 48. Blauwkamp MN, Yu J, Schin MA, Burke KA, Berry MJ, Carlson BA, Brosius FC 3rd, Koenig RJ. Podocyte specific knock out of selenoproteins does not enhance nephropathy in streptozotocin diabetic C57BL/6 mice. BMC Nephrol 9: 7, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 49. Bocci V, Valacchi G. Nrf2 activation as target to implement therapeutic treatments. Front Chem 3: 4, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 50. Boden MJ, Brandon AE, Tid-Ang JD, Preston E, Wilks D, Stuart E, Cleasby ME, Turner N, Cooney GJ, Kraegen EW. Overexpression of manganese superoxide dismutase ameliorates high-fat diet-induced insulin resistance in rat skeletal muscle. Am J Physiol Endocrinol Metab 303: E798–E805, 2012.
    Link | ISI | Google Scholar
  • 51. Bondareva AA, Capecchi MR, Iverson SV, Li Y, Lopez NI, Lucas O, Merrill GF, Prigge JR, Siders AM, Wakamiya M, Wallin SL, Schmidt EE. Effects of thioredoxin reductase-1 deletion on embryogenesis and transcriptome. Free Radic Biol Med 43: 911–923, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 52. Borchert A, Wang CC, Ufer C, Schiebel H, Savaskan NE, Kuhn H. The role of phospholipid hydroperoxide glutathione peroxidase isoforms in murine embryogenesis. J Biol Chem 281: 19655–19664, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 53. Borg J, London J. Copper/zinc superoxide dismutase overexpression promotes survival of cortical neurons exposed to neurotoxins in vitro. J Neurosci Res 70: 180–189, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 54. Bosl MR, Takaku K, Oshima M, Nishimura S, Taketo MM. Early embryonic lethality caused by targeted disruption of the mouse selenocysteine tRNA gene (Trsp). Proc Natl Acad Sci USA 94: 5531–5534, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 55. Bossis G, Melchior F. Regulation of SUMOylation by reversible oxidation of SUMO conjugating enzymes. Mol Cell 21: 349–357, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 56. Boulos S, Meloni BP, Arthur PG, Bojarski C, Knuckey NW. Peroxiredoxin 2 overexpression protects cortical neuronal cultures from ischemic and oxidative injury but not glutamate excitotoxicity, whereas Cu/Zn superoxide dismutase 1 overexpression protects only against oxidative injury. J Neurosci Res 85: 3089–3097, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 57. Brewer AC, Mustafi SB, Murray TV, Rajasekaran NS, Benjamin IJ. Reductive stress linked to small HSPs, G6PD, and Nrf2 pathways in heart disease. Antioxid Redox Signal 18: 1114–1127, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 58. Brezniceanu ML, Liu F, Wei CC, Tran S, Sachetelli S, Zhang SL, Guo DF, Filep JG, Ingelfinger JR, Chan JS. Catalase overexpression attenuates angiotensinogen expression and apoptosis in diabetic mice. Kidney Int 71: 912–923, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 59. Brigelius-Flohe R. Glutathione peroxidases in different stages of carcinogenesis. Biochim Biophys Acta 1790: 1555–1568, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 60. Brigelius-Flohe R. Tissue-specific functions of individual glutathione peroxidases. Free Radic Biol Med 27: 951–965, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 61. Brigelius-Flohe R, Kipp AP. Physiological functions of GPx2 and its role in inflammation-triggered carcinogenesis. Ann NY Acad Sci 1259: 19–25, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 62. Brigelius-Flohe R, Maiorino M. Glutathione peroxidases. Biochim Biophys Acta 1830: 3289–3303, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 63. Britt RD Jr, Velten M, Locy ML, Rogers LK, Tipple TE. The thioredoxin reductase-1 inhibitor aurothioglucose attenuates lung injury and improves survival in a murine model of acute respiratory distress syndrome. Antioxid Redox Signal 20: 2681–2691, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 64. Brown DP, Chin-Sinex H, Nie B, Mendonca MS, Wang M. Targeting superoxide dismutase 1 to overcome cisplatin resistance in human ovarian cancer. Cancer Chemother Pharmacol 63: 723–730, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 65. Brown KA, Didion SP, Andresen JJ, Faraci FM. Effect of aging, MnSOD deficiency, and genetic background on endothelial function: evidence for MnSOD haploinsufficiency. Arterioscler Thromb Vasc Biol 27: 1941–1946, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 66. Brown MR, Miller FJ Jr, Li WG, Ellingson AN, Mozena JD, Chatterjee P, Engelhardt JF, Zwacka RM, Oberley LW, Fang X, Spector AA, Weintraub NL. Overexpression of human catalase inhibits proliferation and promotes apoptosis in vascular smooth muscle cells. Circ Res 85: 524–533, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 67. Bruijn LI, Miller TM, Cleveland DW. Unraveling the mechanisms involved in motor neuron degeneration in ALS. Annu Rev Neurosci 27: 723–749, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 68. Brutsch SH, Wang CC, Li L, Stender H, Neziroglu N, Richter C, Kuhn H, Borchert A. Expression of inactive glutathione peroxidase 4 leads to embryonic lethality, and inactivation of the Alox15 gene does not rescue such knock-in mice. Antioxid Redox Signal 22: 281–293, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 69. Buettner GR. Superoxide dismutase in redox biology: the roles of superoxide and hydrogen peroxide. Anticancer Agents Med Chem 11: 341–346, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 70. Burgoyne JR, Oka S, Ale-Agha N, Eaton P. Hydrogen peroxide sensing and signaling by protein kinases in the cardiovascular system. Antioxid Redox Signal 18: 1042–1052, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 71. Burk RF, Hill KE. Selenoprotein P-expression, functions, and roles in mammals. Biochim Biophys Acta 1790: 1441–1447, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 72. Burk RF, Olson GE, Hill KE, Winfrey VP, Motley AK, Kurokawa S. Maternal-fetal transfer of selenium in the mouse. FASEB J 27: 3249–3256, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 73. Busuttil RA, Garcia AM, Cabrera C, Rodriguez A, Suh Y, Kim WH, Huang TT, Vijg J. Organ-specific increase in mutation accumulation and apoptosis rate in CuZn-superoxide dismutase-deficient mice. Cancer Res 65: 11271–11275, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 74. Calabrese V, Cornelius C, Dinkova-Kostova AT, Iavicoli I, Di Paola R, Koverech A, Cuzzocrea S, Rizzarelli E, Calabrese EJ. Cellular stress responses, hormetic phytochemicals and vitagenes in aging and longevity. Biochim Biophys Acta 1822: 753–783, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 75. Cardoso BR, Ong TP, Jacob-Filho W, Jaluul O, Freitas MI, Cominetti C, Cozzolino SM. Glutathione peroxidase 1 Pro198Leu polymorphism in Brazilian Alzheimer's disease patients: relations to the enzyme activity and to selenium status. J Nutrigenet Nutrigenomics 5: 72–80, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 76. Carlson BA, Novoselov SV, Kumaraswamy E, Lee BJ, Anver MR, Gladyshev VN, Hatfield DL. Specific excision of the selenocysteine tRNA[Ser]Sec (Trsp) gene in mouse liver demonstrates an essential role of selenoproteins in liver function. J Biol Chem 279: 8011–8017, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 77. Carlson BA, Schweizer U, Perella C, Shrimali RK, Feigenbaum L, Shen L, Speransky S, Floss T, Jeong SJ, Watts J, Hoffmann V, Combs GF, Gladyshev VN, Hatfield DL. The selenocysteine tRNA STAF-binding region is essential for adequate selenocysteine tRNA status, selenoprotein expression and early age survival of mice. Biochem J 418: 61–71, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 78. Carlson BA, Xu XM, Gladyshev VN, Hatfield DL. Selective rescue of selenoprotein expression in mice lacking a highly specialized methyl group in selenocysteine tRNA. J Biol Chem 280: 5542–5548, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 79. Carlson BA, Yoo MH, Tobe R, Mueller C, Naranjo-Suarez S, Hoffmann VJ, Gladyshev VN, Hatfield DL. Thioredoxin reductase 1 protects against chemically induced hepatocarcinogenesis via control of cellular redox homeostasis. Carcinogenesis 33: 1806–1813, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 80. Carlson BA, Yoo MH, Tsuji PA, Gladyshev VN, Hatfield DL. Mouse models targeting selenocysteine tRNA expression for elucidating the role of selenoproteins in health and development. Molecules 14: 3509–3527, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 81. Carlsson LM, Jonsson J, Edlund T, Marklund SL. Mice lacking extracellular superoxide dismutase are more sensitive to hyperoxia. Proc Natl Acad Sci USA 92: 6264–6268, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 82. Carlstrom M, Brown RD, Sallstrom J, Larsson E, Zilmer M, Zabihi S, Eriksson UJ, Persson AE. SOD1 deficiency causes salt sensitivity and aggravates hypertension in hydronephrosis. Am J Physiol Regul Integr Comp Physiol 297: R82–R92, 2009.
    Link | ISI | Google Scholar
  • 83. Carlstrom M, Lai EY, Ma Z, Steege A, Patzak A, Eriksson UJ, Lundberg JO, Wilcox CS, Persson AE. Superoxide dismutase 1 limits renal microvascular remodeling and attenuates arteriole and blood pressure responses to angiotensin II via modulation of nitric oxide bioavailability. Hypertension 56: 907–913, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 84. Castellano S. On the unique function of selenocysteine: insights from the evolution of selenoproteins. Biochim Biophys Acta 1790: 1463–1470, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 85. Castello PR, David PS, McClure T, Crook Z, Poyton RO. Mitochondrial cytochrome oxidase produces nitric oxide under hypoxic conditions: implications for oxygen sensing and hypoxic signaling in eukaryotes. Cell Metab 3: 277–287, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 86. Castello PR, Woo DK, Ball K, Wojcik J, Liu L, Poyton RO. Oxygen-regulated isoforms of cytochrome c oxidase have differential effects on its nitric oxide production and on hypoxic signaling. Proc Natl Acad Sci USA 105: 8203–8208, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 87. Ceballos-Picot I, Nicole A, Briand P, Grimber G, Delacourte A, Defossez A, Javoy-Agid F, Lafon M, Blouin JL, Sinet PM. Neuronal-specific expression of human copper-zinc superoxide dismutase gene in transgenic mice: animal model of gene dosage effects in Down's syndrome. Brain Res 552: 198–214, 1991.
    Crossref | PubMed | ISI | Google Scholar
  • 88. Cebula M, Moolla N, Capovilla A, Arner ES. The rare TXNRD1_v3 (“v3”) splice variant of human thioredoxin reductase 1 protein is targeted to membrane rafts by N-acylation and induces filopodia independently of its redox active site integrity. J Biol Chem 288: 10002–10011, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 89. Cebula M, Schmidt EE, Arner ES. TrxR1 as a potent regulator of the Nrf2-Keap1 response system. Antioxid Redox Signal 23: 823–853, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 90. Chae HZ, Kim HJ, Kang SW, Rhee SG. Characterization of three isoforms of mammalian peroxiredoxin that reduce peroxides in the presence of thioredoxin. Diabetes Res Clin Pract 45: 101–112, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 91. Chakravarti R, Stuehr DJ. Thioredoxin-1 regulates cellular heme insertion by controlling S-nitrosation of glyceraldehyde-3-phosphate dehydrogenase. J Biol Chem 287: 16179–16186, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 92. Chan PH, Chu L, Chen SF, Carlson EJ, Epstein CJ. Reduced neurotoxicity in transgenic mice overexpressing human copper-zinc-superoxide dismutase. Stroke 21: III80–82, 1990.
    PubMed | ISI | Google Scholar
  • 93. Chan PH, Yang GY, Chen SF, Carlson E, Epstein CJ. Cold-induced brain edema and infarction are reduced in transgenic mice overexpressing CuZn-superoxide dismutase. Ann Neurol 29: 482–486, 1991.
    Crossref | PubMed | ISI | Google Scholar
  • 94. Chance B, Sies H, Boveris A. Hydroperoxide metabolism in mammalian organs. Physiol Rev 59: 527–605, 1979.
    Link | ISI | Google Scholar
  • 95. Chang EY, Son SK, Ko HS, Baek SH, Kim JH, Kim JR. Induction of apoptosis by the overexpression of an alternative splicing variant of mitochondrial thioredoxin reductase. Free Radic Biol Med 39: 1666–1675, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 96. Chen CA, Wang TY, Varadharaj S, Reyes LA, Hemann C, Talukder MA, Chen YR, Druhan LJ, Zweier JL. S-glutathionylation uncouples eNOS and regulates its cellular and vascular function. Nature 468: 1115–1118, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 97. Chen EP, Bittner HB, Davis RD, Folz RJ, Van Trigt P. Extracellular superoxide dismutase transgene overexpression preserves postischemic myocardial function in isolated murine hearts. Circulation 94: II412–417, 1996.
    PubMed | ISI | Google Scholar
  • 98. Chen EP, Bittner HB, Davis RD, Van Trigt P, Folz RJ. Physiologic effects of extracellular superoxide dismutase transgene overexpression on myocardial function after ischemia and reperfusion injury. J Thorac Cardiovasc Surg 115: 450–459, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 99. Chen H, Li X, Epstein PN. MnSOD and catalase transgenes demonstrate that protection of islets from oxidative stress does not alter cytokine toxicity. Diabetes 54: 1437–1446, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 100. Chen H, Yoshioka H, Kim GS, Jung JE, Okami N, Sakata H, Maier CM, Narasimhan P, Goeders CE, Chan PH. Oxidative stress in ischemic brain damage: mechanisms of cell death and potential molecular targets for neuroprotection. Antioxid Redox Signal 14: 1505–1517, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 101. Chen L, Na R, Gu M, Richardson A, Ran Q. Lipid peroxidation up-regulates BACE1 expression in vivo: a possible early event of amyloidogenesis in Alzheimer's disease. J Neurochem 107: 197–207, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 102. Chen L, Na R, Gu M, Salmon AB, Liu Y, Liang H, Qi W, Van Remmen H, Richardson A, Ran Q. Reduction of mitochondrial H2O2 by overexpressing peroxiredoxin 3 improves glucose tolerance in mice. Aging Cell 7: 866–878, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 103. Chen X, Mele J, Giese H, Van Remmen H, Dolle ME, Steinhelper M, Richardson A, Vijg J. A strategy for the ubiquitous overexpression of human catalase and CuZn superoxide dismutase genes in transgenic mice. Mech Ageing Dev 124: 219–227, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 104. Chen Y, Chan PH, Swanson RA. Astrocytes overexpressing Cu,Zn superoxide dismutase have increased resistance to oxidative injury. Glia 33: 343–347, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 105. Chen Y, Yu A, Saari JT, Kang YJ. Repression of hypoxia-reoxygenation injury in the catalase-overexpressing heart of transgenic mice. Proc Soc Exp Biol Med 216: 112–116, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 106. Chen Z, Oberley TD, Ho Y, Chua CC, Siu B, Hamdy RC, Epstein CJ, Chua BH. Overexpression of CuZnSOD in coronary vascular cells attenuates myocardial ischemia/reperfusion injury. Free Radic Biol Med 29: 589–596, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 107. Chen Z, Siu B, Ho YS, Vincent R, Chua CC, Hamdy RC, Chua BH. Overexpression of MnSOD protects against myocardial ischemia/reperfusion injury in transgenic mice. J Mol Cell Cardiol 30: 2281–2289, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 108. Cheng W, Fu YX, Porres JM, Ross DA, Lei XG. Selenium-dependent cellular glutathione peroxidase protects mice against a pro-oxidant-induced oxidation of NADPH, NADH, lipids, and protein. FASEB J 13: 1467–1475, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 109. Cheng WH, Ho YS, Ross DA, Han Y, Combs GF Jr, Lei XG. Overexpression of cellular glutathione peroxidase does not affect expression of plasma glutathione peroxidase or phospholipid hydroperoxide glutathione peroxidase in mice offered diets adequate or deficient in selenium. J Nutr 127: 675–680, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 110. Cheng WH, Ho YS, Ross DA, Valentine BA, Combs GF, Lei XG. Cellular glutathione peroxidase knockout mice express normal levels of selenium-dependent plasma and phospholipid hydroperoxide glutathione peroxidases in various tissues. J Nutr 127: 1445–1450, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 111. Cheng WH, Ho YS, Valentine BA, Ross DA, Combs GF Jr, Lei XG. Cellular glutathione peroxidase is the mediator of body selenium to protect against paraquat lethality in transgenic mice. J Nutr 128: 1070–1076, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 112. Cheng WH, Quimby FW, Lei XG. Impacts of glutathione peroxidase-1 knockout on the protection by injected selenium against the pro-oxidant-induced liver aponecrosis and signaling in selenium-deficient mice. Free Radic Biol Med 34: 918–927, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 113. Cheng WH, Valentine BA, Lei XG. High levels of dietary vitamin E do not replace cellular glutathione peroxidase in protecting mice from acute oxidative stress. J Nutr 129: 1951–1957, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 114. Cheng WH, Zheng X, Quimby FR, Roneker CA, Lei XG. Low levels of glutathione peroxidase 1 activity in selenium-deficient mouse liver affect c-Jun N-terminal kinase activation and p53 phosphorylation on Ser-15 in pro-oxidant-induced aponecrosis. Biochem J 370: 927–934, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 115. Chew P, Yuen DY, Stefanovic N, Pete J, Coughlan MT, Jandeleit-Dahm KA, Thomas MC, Rosenfeldt F, Cooper ME, de Haan JB. Antiatherosclerotic and renoprotective effects of ebselen in the diabetic apolipoprotein E/GPx1-double knockout mouse. Diabetes 59: 3198–3207, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 116. Cho HY, Jedlicka AE, Reddy SP, Zhang LY, Kensler TW, Kleeberger SR. Linkage analysis of susceptibility to hyperoxia. Nrf2 is a candidate gene. Am J Respir Cell Mol Biol 26: 42–51, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 117. Chu FF, Doroshow JH, Esworthy RS. Expression, characterization, and tissue distribution of a new cellular selenium-dependent glutathione peroxidase, GSHPx-GI. J Biol Chem 268: 2571–2576, 1993.
    PubMed | ISI | Google Scholar
  • 118. Chu FF, Esworthy RS, Chu PG, Longmate JA, Huycke MM, Wilczynski S, Doroshow JH. Bacteria-induced intestinal cancer in mice with disrupted Gpx1 and Gpx2 genes. Cancer Res 64: 962–968, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 119. Clark RF, Strukle E, Williams SR, Manoguerra AS. Selenium poisoning from a nutritional supplement. JAMA 275: 1087–1088, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 120. Coling DE, Yu KC, Somand D, Satar B, Bai U, Huang TT, Seidman MD, Epstein CJ, Mhatre AN, Lalwani AK. Effect of SOD1 overexpression on age- and noise-related hearing loss. Free Radic Biol Med 34: 873–880, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 121. Cominetti C, de Bortoli MC, Purgatto E, Ong TP, Moreno FS, Garrido AB Jr, Cozzolino SM. Associations between glutathione peroxidase-1 Pro198Leu polymorphism, selenium status, and DNA damage levels in obese women after consumption of Brazil nuts. Nutrition 27: 891–896, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 122. Conrad M. Transgenic mouse models for the vital selenoenzymes cytosolic thioredoxin reductase, mitochondrial thioredoxin reductase and glutathione peroxidase 4. Biochim Biophys Acta 1790: 1575–1585, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 123. Conrad M, Jakupoglu C, Moreno SG, Lippl S, Banjac A, Schneider M, Beck H, Hatzopoulos AK, Just U, Sinowatz F, Schmahl W, Chien KR, Wurst W, Bornkamm GW, Brielmeier M. Essential role for mitochondrial thioredoxin reductase in hematopoiesis, heart development, and heart function. Mol Cell Biol 24: 9414–9423, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 124. Conrad M, Schweizer U. Unveiling the molecular mechanisms behind selenium-related diseases through knockout mouse studies. Antioxid Redox Signal 12: 851–865, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 125. Cooke CL, Davidge ST. Endothelial-dependent vasodilation is reduced in mesenteric arteries from superoxide dismutase knockout mice. Cardiovasc Res 60: 635–642, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 126. Corniola R, Zou Y, Leu D, Fike JR, Huang TT. Paradoxical relationship between Mn superoxide dismutase deficiency and radiation-induced cognitive defects. PLoS One 7: e49367, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 127. Corson LB, Strain J, Culotta VC, Cleveland DW. Chaperone-facilitated copper binding is a property common to several classes of familial amyotrophic lateral sclerosis-linked superoxide dismutase mutants. Proc Natl Acad Sci USA 95: 6361–6366, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 128. Crack PJ, Cimdins K, Ali U, Hertzog PJ, Iannello RC. Lack of glutathione peroxidase-1 exacerbates Abeta-mediated neurotoxicity in cortical neurons. J Neural Transm 113: 645–657, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 129. Craven PA, Melhem MF, Phillips SL, DeRubertis FR. Overexpression of Cu2+/Zn2+ superoxide dismutase protects against early diabetic glomerular injury in transgenic mice. Diabetes 50: 2114–2125, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 130. Crawford A, Fassett RG, Geraghty DP, Kunde DA, Ball MJ, Robertson IK, Coombes JS. Relationships between single nucleotide polymorphisms of antioxidant enzymes and disease. Gene 501: 89–103, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 131. Crosti N, Serra A, Rigo A, Viglino P. Dosage effect of SOD-A gene in 21-trisomic cells. Hum Genet 31: 197–202, 1976.
    Crossref | PubMed | ISI | Google Scholar
  • 132. Culotta VC, Joh HD, Lin SJ, Slekar KH, Strain J. A physiological role for Saccharomyces cerevisiae copper/zinc superoxide dismutase in copper buffering. J Biol Chem 270: 29991–29997, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 133. Curry-McCoy TV, Osna NA, Nanji AA, Donohue TM Jr. Chronic ethanol consumption results in atypical liver injury in copper/zinc superoxide dismutase deficient mice. Alcohol Clin Exp Res 34: 251–261, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 134. Czernichow S, Couthouis A, Bertrais S, Vergnaud AC, Dauchet L, Galan P, Hercberg S. Antioxidant supplementation does not affect fasting plasma glucose in the Supplementation with Antioxidant Vitamins and Minerals (SUVI MAX) study in France: association with dietary intake and plasma concentrations. Am J Clin Nutr 84: 395–399, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 135. D'Autreaux B, Toledano MB. ROS as signalling molecules: mechanisms that generate specificity in ROS homeostasis. Nat Rev Mol Cell Biol 8: 813–824, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 136. Dabkowski ER, Williamson CL, Hollander JM. Mitochondria-specific transgenic overexpression of phospholipid hydroperoxide glutathione peroxidase (GPx4) attenuates ischemia/reperfusion-associated cardiac dysfunction. Free Radic Biol Med 45: 855–865, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 137. Dai DF, Santana LF, Vermulst M, Tomazela DM, Emond MJ, MacCoss MJ, Gollahon K, Martin GM, Loeb LA, Ladiges WC, Rabinovitch PS. Overexpression of catalase targeted to mitochondria attenuates murine cardiac aging. Circulation 119: 2789–2797, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 138. Daiber A, Oelze M, Sulyok S, Coldewey M, Schulz E, Treiber N, Hink U, Mulsch A, Scharffetter-Kochanek K, Munzel T. Heterozygous deficiency of manganese superoxide dismutase in mice (Mn-SOD+/−): a novel approach to assess the role of oxidative stress for the development of nitrate tolerance. Mol Pharmacol 68: 579–588, 2005.
    PubMed | ISI | Google Scholar
  • 139. Damdimopoulos AE, Miranda-Vizuete A, Treuter E, Gustafsson JA, Spyrou G. An alternative splicing variant of the selenoprotein thioredoxin reductase is a modulator of estrogen signaling. J Biol Chem 279: 38721–38729, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 140. Das KC. Thioredoxin-deficient mice, a novel phenotype sensitive to ambient air and hypersensitive to hyperoxia-induced lung injury. Am J Physiol Lung Cell Mol Physiol 308: L429–L442, 2015.
    Link | ISI | Google Scholar
  • 141. De Haan JB, Bladier C, Griffiths P, Kelner M, O'Shea RD, Cheung NS, Bronson RT, Silvestro MJ, Wild S, Zheng SS, Beart PM, Hertzog PJ, Kola I. Mice with a homozygous null mutation for the most abundant glutathione peroxidase, Gpx1, show increased susceptibility to the oxidative stress-inducing agents paraquat and hydrogen peroxide. J Biol Chem 273: 22528–22536, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 142. De Haan JB, Cristiano F, Iannello R, Bladier C, Kelner MJ, Kola I. Elevation in the ratio of Cu/Zn-superoxide dismutase to glutathione peroxidase activity induces features of cellular senescence and this effect is mediated by hydrogen peroxide. Hum Mol Genet 5: 283–292, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 143. De La Torre R, Casado A, Lopez-Fernandez E, Carrascosa D, Ramirez V, Saez J. Overexpression of copper-zinc superoxide dismutase in trisomy 21. Experientia 52: 871–873, 1996.
    Crossref | PubMed | Google Scholar
  • 144. De Vos S, Epstein CJ, Carlson E, Cho SK, Koeffler HP. Transgenic mice overexpressing human copper/zinc-superoxide dismutase (Cu/Zn SOD) are not resistant to endotoxic shock. Biochem Biophys Res Commun 208: 523–531, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 145. DeNicola GM, Karreth FA, Humpton TJ, Gopinathan A, Wei C, Frese K, Mangal D, Yu KH, Yeo CJ, Calhoun ES, Scrimieri F, Winter JM, Hruban RH, Iacobuzio-Donahue C, Kern SE, Blair IA, Tuveson DA. Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 475: 106–109, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 146. Denu JM, Tanner KG. Specific and reversible inactivation of protein tyrosine phosphatases by hydrogen peroxide: evidence for a sulfenic acid intermediate and implications for redox regulation. Biochemistry 37: 5633–5642, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 147. DeRubertis FR, Craven PA, Melhem MF. Acceleration of diabetic renal injury in the superoxide dismutase knockout mouse: effects of tempol. Metabolism 56: 1256–1264, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 148. DeRubertis FR, Craven PA, Melhem MF, Salah EM. Attenuation of renal injury in db/db mice overexpressing superoxide dismutase: evidence for reduced superoxide-nitric oxide interaction. Diabetes 53: 762–768, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 149. Dhar SK, St Clair DK. Manganese superoxide dismutase regulation and cancer. Free Radic Biol Med 52: 2209–2222, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 150. Di Cosmo C, McLellan N, Liao XH, Khanna KK, Weiss RE, Papp L, Refetoff S. Clinical and molecular characterization of a novel selenocysteine insertion sequence-binding protein 2 (SBP2) gene mutation (R128X). J Clin Endocrinol Metab 94: 4003–4009, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 151. Didion SP, Kinzenbaw DA, Schrader LI, Faraci FM. Heterozygous CuZn superoxide dismutase deficiency produces a vascular phenotype with aging. Hypertension 48: 1072–1079, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 152. Didion SP, Ryan MJ, Didion LA, Fegan PE, Sigmund CD, Faraci FM. Increased superoxide and vascular dysfunction in CuZnSOD-deficient mice. Circ Res 91: 938–944, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 153. Dimayuga FO, Wang C, Clark JM, Dimayuga ER, Dimayuga VM, Bruce-Keller AJ. SOD1 overexpression alters ROS production and reduces neurotoxic inflammatory signaling in microglial cells. J Neuroimmunol 182: 89–99, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 154. Ding H, Schwarz DS, Keene A, Affar el B, Fenton L, Xia X, Shi Y, Zamore PD, Xu Z. Selective silencing by RNAi of a dominant allele that causes amyotrophic lateral sclerosis. Aging Cell 2: 209–217, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 155. Ding Y, Yamada S, Wang KY, Shimajiri S, Guo X, Tanimoto A, Murata Y, Kitajima S, Watanabe T, Izumi H, Kohno K, Sasaguri Y. Overexpression of peroxiredoxin 4 protects against high-dose streptozotocin-induced diabetes by suppressing oxidative stress and cytokines in transgenic mice. Antioxid Redox Signal 13: 1477–1490, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 156. Dinkova-Kostova AT. Chemoprotection against cancer by isothiocyanates: a focus on the animal models and the protective mechanisms. Top Curr Chem 329: 179–201, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 157. Dinkova-Kostova AT, Holtzclaw WD, Cole RN, Itoh K, Wakabayashi N, Katoh Y, Yamamoto M, Talalay P. Direct evidence that sulfhydryl groups of Keap1 are the sensors regulating induction of phase 2 enzymes that protect against carcinogens and oxidants. Proc Natl Acad Sci USA 99: 11908–11913, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 158. Dirmeier R, O'Brien KM, Engle M, Dodd A, Spears E, Poyton RO. Exposure of yeast cells to anoxia induces transient oxidative stress. Implications for the induction of hypoxic genes. J Biol Chem 277: 34773–34784, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 159. Diwadkar-Navsariwala V, Prins GS, Swanson SM, Birch LA, Ray VH, Hedayat S, Lantvit DL, Diamond AM. Selenoprotein deficiency accelerates prostate carcinogenesis in a transgenic model. Proc Natl Acad Sci USA 103: 8179–8184, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 160. Dong F, Fang CX, Yang X, Zhang X, Lopez FL, Ren J. Cardiac overexpression of catalase rescues cardiac contractile dysfunction induced by insulin resistance: role of oxidative stress, protein carbonyl formation and insulin sensitivity. Diabetologia 49: 1421–1433, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 161. Downey CM, Horton CR, Carlson BA, Parsons TE, Hatfield DL, Hallgrimsson B, Jirik FR. Osteo-chondroprogenitor-specific deletion of the selenocysteine tRNA gene, Trsp, leads to chondronecrosis and abnormal skeletal development: a putative model for Kashin-Beck disease. PLoS Genet 5: e1000616, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 162. Du Y, Zhang H, Lu J, Holmgren A. Glutathione and glutaredoxin act as a backup of human thioredoxin reductase 1 to reduce thioredoxin 1 preventing cell death by aurothioglucose. J Biol Chem 287: 38210–38219, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 163. Du Y, Zhang H, Montano S, Hegestam J, Ekberg NR, Holmgren A, Brismar K, Ungerstedt JS. Plasma glutaredoxin activity in healthy subjects and patients with abnormal glucose levels or overt type 2 diabetes. Acta Diabetol 51: 225–232, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 164. Dumont M, Wille E, Stack C, Calingasan NY, Beal MF, Lin MT. Reduction of oxidative stress, amyloid deposition, and memory deficit by manganese superoxide dismutase overexpression in a transgenic mouse model of Alzheimer's disease. FASEB J 23: 2459–2466, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 165. Duong C, Seow HJ, Bozinovski S, Crack PJ, Anderson GP, Vlahos R. Glutathione peroxidase-1 protects against cigarette smoke-induced lung inflammation in mice. Am J Physiol Lung Cell Mol Physiol 299: L425–L433, 2010.
    Link | ISI | Google Scholar
  • 166. Eaton JW, Ma M. Acatalsaemia In: The Metabolic Bases of Inherited Disease, edited by Scriver C, Beudet A, Sly W, Valle DL. New York: McGraw-Hill, 1995, p. 2371–2383.
    Google Scholar
  • 167. Elchuri S, Oberley TD, Qi W, Eisenstein RS, Jackson Roberts L, Van Remmen H, Epstein CJ, Huang TT. CuZnSOD deficiency leads to persistent and widespread oxidative damage and hepatocarcinogenesis later in life. Oncogene 24: 367–380, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 168. Elroy-Stein O, Bernstein Y, Groner Y. Overproduction of human Cu/Zn-superoxide dismutase in transfected cells: extenuation of paraquat-mediated cytotoxicity and enhancement of lipid peroxidation. EMBO J 5: 615–622, 1986.
    Crossref | PubMed | ISI | Google Scholar
  • 169. Endo H, Nito C, Kamada H, Yu F, Chan PH. Reduction in oxidative stress by superoxide dismutase overexpression attenuates acute brain injury after subarachnoid hemorrhage via activation of Akt/glycogen synthase kinase-3beta survival signaling. J Cereb Blood Flow Metab 27: 975–982, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 170. Epstein CJ, Avraham KB, Lovett M, Smith S, Elroy-Stein O, Rotman G, Bry C, Groner Y. Transgenic mice with increased Cu/Zn-superoxide dismutase activity: animal model of dosage effects in Down syndrome. Proc Natl Acad Sci USA 84: 8044–8048, 1987.
    Crossref | PubMed | ISI | Google Scholar
  • 171. Eriksson S, Prigge JR, Talago EA, Arner ES, Schmidt EE. Dietary methionine can sustain cytosolic redox homeostasis in the mouse liver. Nat Commun 6: 6479, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 172. Esposito LA, Kokoszka JE, Waymire KG, Cottrell B, MacGregor GR, Wallace DC. Mitochondrial oxidative stress in mice lacking the glutathione peroxidase-1 gene. Free Radic Biol Med 28: 754–766, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 173. Estevez AG, Crow JP, Sampson JB, Reiter C, Zhuang Y, Richardson GJ, Tarpey MM, Barbeito L, Beckman JS. Induction of nitric oxide-dependent apoptosis in motor neurons by zinc-deficient superoxide dismutase. Science 286: 2498–2500, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 174. Esworthy RS, Aranda R, Martin MG, Doroshow JH, Binder SW, Chu FF. Mice with combined disruption of Gpx1 and Gpx2 genes have colitis. Am J Physiol Gastrointest Liver Physiol 281: G848–G855, 2001.
    Link | ISI | Google Scholar
  • 175. Esworthy RS, Binder SW, Doroshow JH, Chu FF. Microflora trigger colitis in mice deficient in selenium-dependent glutathione peroxidase and induce Gpx2 gene expression. Biol Chem 384: 597–607, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 176. Esworthy RS, Mann JR, Sam M, Chu FF. Low glutathione peroxidase activity in Gpx1 knockout mice protects jejunum crypts from gamma-irradiation damage. Am J Physiol Gastrointest Liver Physiol 279: G426–G436, 2000.
    Link | ISI | Google Scholar
  • 177. Esworthy RS, Swiderek KM, Ho YS, Chu FF. Selenium-dependent glutathione peroxidase-GI is a major glutathione peroxidase activity in the mucosal epithelium of rodent intestine. Biochim Biophys Acta 1381: 213–226, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 178. Faraci FM, Didion SP. Vascular protection: superoxide dismutase isoforms in the vessel wall. Arterioscler Thromb Vasc Biol 24: 1367–1373, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 179. Faraci FM, Modrick ML, Lynch CM, Didion LA, Fegan PE, Didion SP. Selective cerebral vascular dysfunction in Mn-SOD-deficient mice. J Appl Physiol 100: 2089–2093, 2006.
    Link | ISI | Google Scholar
  • 180. Fernandes AP, Holmgren A. Glutaredoxins: glutathione-dependent redox enzymes with functions far beyond a simple thioredoxin backup system. Antioxid Redox Signal 6: 63–74, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 181. Ferrer-Sueta G, Manta B, Botti H, Radi R, Trujillo M, Denicola A. Factors affecting protein thiol reactivity and specificity in peroxide reduction. Chem Res Toxicol 24: 434–450, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 182. Finley JW, Kong AN, Hintze KJ, Jeffery EH, Ji LL, Lei XG. Antioxidants in foods: state of the science important to the food industry. J Agric Food Chem 59: 6837–6846, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 183. Fishman K, Baure J, Zou Y, Huang TT, Andres-Mach M, Rola R, Suarez T, Acharya M, Limoli CL, Lamborn KR, Fike JR. Radiation-induced reductions in neurogenesis are ameliorated in mice deficient in CuZnSOD or MnSOD. Free Radic Biol Med 47: 1459–1467, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 184. Flohe L, Loschen G, Gunzler WA, Eichele E. Glutathione peroxidase. V. The kinetic mechanism. Hoppe-Seylers Z Physiol Chem 353: 987–999, 1972.
    Crossref | PubMed | Google Scholar
  • 185. Flohe L, Toppo S, Cozza G, Ursini F. A comparison of thiol peroxidase mechanisms. Antioxid Redox Signal 15: 763–780, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 186. Florian S, Krehl S, Loewinger M, Kipp A, Banning A, Esworthy S, Chu FF, Brigelius-Flohe R. Loss of GPx2 increases apoptosis, mitosis, and GPx1 expression in the intestine of mice. Free Radic Biol Med 49: 1694–1702, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 187. Florian S, Wingler K, Schmehl K, Jacobasch G, Kreuzer OJ, Meyerhof W, Brigelius-Flohe R. Cellular and subcellular localization of gastrointestinal glutathione peroxidase in normal and malignant human intestinal tissue. Free Radic Res 35: 655–663, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 188. Flynn JM, Choi SW, Day NU, Gerencser AA, Hubbard A, Melov S. Impaired spare respiratory capacity in cortical synaptosomes from Sod2 null mice. Free Radic Biol Med 50: 866–873, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 189. Folz RJ, Abushamaa AM, Suliman HB. Extracellular superoxide dismutase in the airways of transgenic mice reduces inflammation and attenuates lung toxicity following hyperoxia. J Clin Invest 103: 1055–1066, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 190. Fomenko DE, Novoselov SV, Natarajan SK, Lee BC, Koc A, Carlson BA, Lee TH, Kim HY, Hatfield DL, Gladyshev VN. MsrB1 (methionine-R-sulfoxide reductase 1) knock-out mice: roles of MsrB1 in redox regulation and identification of a novel selenoprotein form. J Biol Chem 284: 5986–5993, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 191. Forgione MA, Weiss N, Heydrick S, Cap A, Klings ES, Bierl C, Eberhardt RT, Farber HW, Loscalzo J. Cellular glutathione peroxidase deficiency and endothelial dysfunction. Am J Physiol Heart Circ Physiol 282: H1255–H1261, 2002.
    Link | ISI | Google Scholar
  • 192. Forman HJ, Fukuto JM, Torres M. Redox signaling: thiol chemistry defines which reactive oxygen and nitrogen species can act as second messengers. Am J Physiol Cell Physiol 287: C246–C256, 2004.
    Link | ISI | Google Scholar
  • 193. Forman HJ, Maiorino M, Ursini F. Signaling functions of reactive oxygen species. Biochemistry 49: 835–842, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 194. Friedmann Angeli JP, Schneider M, Proneth B, Tyurina YY, Tyurin VA, Hammond VJ, Herbach N, Aichler M, Walch A, Eggenhofer E, Basavarajappa D, Radmark O, Kobayashi S, Seibt T, Beck H, Neff F, Esposito I, Wanke R, Forster H, Yefremova O, Heinrichmeyer M, Bornkamm GW, Geissler EK, Thomas SB, Stockwell BR, O'Donnell VB, Kagan VE, Schick JA, Conrad M. Inactivation of the ferroptosis regulator Gpx4 triggers acute renal failure in mice. Nat Cell Biol 16: 1180–1191, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 195. Fu Y, Cheng WH, Porres JM, Ross DA, Lei XG. Knockout of cellular glutathione peroxidase gene renders mice susceptible to diquat-induced oxidative stress. Free Radic Biol Med 27: 605–611, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 196. Fu Y, Cheng WH, Ross DA, Lei X. Cellular glutathione peroxidase protects mice against lethal oxidative stress induced by various doses of diquat. Proc Soc Exp Biol Med 222: 164–169, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 197. Fu Y, McCormick CC, Roneker C, Lei XG. Lipopolysaccharide and interferon-gamma-induced nitric oxide production and protein oxidation in mouse peritoneal macrophages are affected by glutathione peroxidase-1 gene knockout. Free Radic Biol Med 31: 450–459, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 198. Fu Y, Porres JM, Lei XG. Comparative impacts of glutathione peroxidase-1 gene knockout on oxidative stress induced by reactive oxygen and nitrogen species in mouse hepatocytes. Biochem J 359: 687–695, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 199. Fu Y, Sies H, Lei XG. Opposite roles of selenium-dependent glutathione peroxidase-1 in superoxide generator diquat- and peroxynitrite-induced apoptosis and signaling. J Biol Chem 276: 43004–43009, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 200. Fujimoto K, Kumagai K, Ito K, Arakawa S, Ando Y, Oda S, Yamoto T, Manabe S. Sensitivity of liver injury in heterozygous Sod2 knockout mice treated with troglitazone or acetaminophen. Toxicol Pathol 37: 193–200, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 201. Gahtan E, Auerbach JM, Groner Y, Segal M. Reversible impairment of long-term potentiation in transgenic Cu/Zn-SOD mice. Eur J Neurosci 10: 538–544, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 202. Galasso G, Schiekofer S, Sato K, Shibata R, Handy DE, Ouchi N, Leopold JA, Loscalzo J, Walsh K. Impaired angiogenesis in glutathione peroxidase-1-deficient mice is associated with endothelial progenitor cell dysfunction. Circ Res 98: 254–261, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 203. Galbiati M, Crippa V, Rusmini P, Cristofani R, Cicardi ME, Giorgetti E, Onesto E, Messi E, Poletti A. ALS-related misfolded protein management in motor neurons and muscle cells. Neurochem Int 79: 70–78, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 204. Gan L, Yang XL, Liu Q, Xu HB. Inhibitory effects of thioredoxin reductase antisense RNA on the growth of human hepatocellular carcinoma cells. J Cell Biochem 96: 653–664, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 205. Gao F, Kinnula VL, Myllarniemi M, Oury TD. Extracellular superoxide dismutase in pulmonary fibrosis. Antioxid Redox Signal 10: 343–354, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 206. Gao J, Xiong Y, Ho YS, Liu X, Chua CC, Xu X, Wang H, Hamdy R, Chua BH. Glutathione peroxidase 1-deficient mice are more susceptible to doxorubicin-induced cardiotoxicity. Biochim Biophys Acta 1783: 2020–2029, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 207. Garratt M, Bathgate R, de Graaf S, Brooks RC. Copper-zinc superoxide dismutase deficiency impairs sperm motility and in vivo fertility. Reproduction 146: 297–304, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 208. Ge W, Zhang Y, Han X, Ren J. Cardiac-specific overexpression of catalase attenuates paraquat-induced myocardial geometric and contractile alteration: role of ER stress. Free Radic Biol Med 49: 2068–2077, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 209. Geisberger R, Kiermayer C, Homig C, Conrad M, Schmidt J, Zimber-Strobl U, Brielmeier M. B- and T-cell-specific inactivation of thioredoxin reductase 2 does not impair lymphocyte development and maintenance. Biol Chem 388: 1083–1090, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 210. Geismann C, Arlt A, Sebens S, Schafer H. Cytoprotection “gone astray”: Nrf2 and its role in cancer. Onco Targets Ther 7: 1497–1518, 2014.
    PubMed | ISI | Google Scholar
  • 211. Gerashchenko MV, Su D, Gladyshev VN. CUG start codon generates thioredoxin/glutathione reductase isoforms in mouse testes. J Biol Chem 285: 4595–4602, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 212. Ghio AJ, Suliman HB, Carter JD, Abushamaa AM, Folz RJ. Overexpression of extracellular superoxide dismutase decreases lung injury after exposure to oil fly ash. Am J Physiol Lung Cell Mol Physiol 283: L211–L218, 2002.
    Link | ISI | Google Scholar
  • 213. Gil-Bea F, Akterin S, Persson T, Mateos L, Sandebring A, Avila-Carino J, Gutierrez-Rodriguez A, Sundstrom E, Holmgren A, Winblad B, Cedazo-Minguez A. Thioredoxin-80 is a product of alpha-secretase cleavage that inhibits amyloid-beta aggregation and is decreased in Alzheimer's disease brain. EMBO Mol Med 4: 1097–1111, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 214. Giudice A, Montella M. Activation of the Nrf2-ARE signaling pathway: a promising strategy in cancer prevention. Bioessays 28: 169–181, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 215. Gladyshev VN, Jeang KT, Stadtman TC. Selenocysteine, identified as the penultimate C-terminal residue in human T-cell thioredoxin reductase, corresponds to TGA in the human placental gene. Proc Natl Acad Sci USA 93: 6146–6151, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 216. Glorieux C, Dejeans N, Sid B, Beck R, Calderon PB, Verrax J. Catalase overexpression in mammary cancer cells leads to a less aggressive phenotype and an altered response to chemotherapy. Biochem Pharmacol 82: 1384–1390, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 217. Gluck MR, Jayatilleke E, Shaw S, Rowan AJ, Haroutunian V. CNS oxidative stress associated with the kainic acid rodent model of experimental epilepsy. Epilepsy Res 39: 63–71, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 218. Go YM, Orr M, Jones DP. Increased nuclear thioredoxin-1 potentiates cadmium-induced cytotoxicity. Toxicol Sci 131: 84–94, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 219. Godin N, Liu F, Lau GJ, Brezniceanu ML, Chenier I, Filep JG, Ingelfinger JR, Zhang SL, Chan JS. Catalase overexpression prevents hypertension and tubular apoptosis in angiotensinogen transgenic mice. Kidney Int 77: 1086–1097, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 220. Golenser J, Peled-Kamar M, Schwartz E, Friedman I, Groner Y, Pollack Y. Transgenic mice with elevated level of CuZnSOD are highly susceptible to malaria infection. Free Radic Biol Med 24: 1504–1510, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 221. Goodman M, Bostick RM, Kucuk O, Jones DP. Clinical trials of antioxidants as cancer prevention agents: past, present, and future. Free Radic Biol Med 51: 1068–1084, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 222. Goth L. A novel catalase mutation (a G insertion in exon 2) causes the type B of the Hungarian acatalasemia. Clin Chim Acta 311: 161–163, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 223. Goth L, Nagy T. Acatalasemia and diabetes mellitus. Arch Biochem Biophys 525: 195–200, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 224. Goth L, Nagy T. Inherited catalase deficiency: is it benign or a factor in various age related disorders? Mutat Res 753: 147–154, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 225. Goth L, Shemirani A, Kalmar T. A novel catalase mutation (a GA insertion) causes the Hungarian type of acatalasemia. Blood Cells Mol Dis 26: 151–154, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 226. Goto H, Nishikawa T, Sonoda K, Kondo T, Kukidome D, Fujisawa K, Yamashiro T, Motoshima H, Matsumura T, Tsuruzoe K, Araki E. Endothelial MnSOD overexpression prevents retinal VEGF expression in diabetic mice. Biochem Biophys Res Commun 366: 814–820, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 227. Grek CL, Zhang J, Manevich Y, Townsend DM, Tew KD. Causes and consequences of cysteine S-glutathionylation. J Biol Chem 288: 26497–26504, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 228. Groleau J, Dussault S, Turgeon J, Haddad P, Rivard A. Accelerated vascular aging in CuZnSOD-deficient mice: impact on EPC function and reparative neovascularization. PLoS One 6: e23308, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 229. Guo S, Dai C, Guo M, Taylor B, Harmon JS, Sander M, Robertson RP, Powers AC, Stein R. Inactivation of specific beta cell transcription factors in type 2 diabetes. J Clin Invest 123: 3305–3316, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 230. Gurgul E, Lortz S, Tiedge M, Jorns A, Lenzen S. Mitochondrial catalase overexpression protects insulin-producing cells against toxicity of reactive oxygen species and proinflammatory cytokines. Diabetes 53: 2271–2280, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 231. Gurney ME, Pu H, Chiu AY, Dal Canto MC, Polchow CY, Alexander DD, Caliendo J, Hentati A, Kwon YW, Deng HX. Motor neuron degeneration in mice that express a human Cu,Zn superoxide dismutase mutation. Science 264: 1772–1775, 1994.
    Crossref | PubMed | ISI | Google Scholar
  • 232. Hagay ZJ, Weiss Y, Zusman I, Peled-Kamar M, Reece EA, Eriksson UJ, Groner Y. Prevention of diabetes-associated embryopathy by overexpression of the free radical scavenger copper zinc superoxide dismutase in transgenic mouse embryos. Am J Obstet Gynecol 173: 1036–1041, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 233. Halliwell B. The antioxidant paradox: less paradoxical now? Br J Clin Pharmacol 75: 637–644, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 234. Halliwell B. Free radicals and antioxidants: updating a personal view. Nutr Rev 70: 257–265, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 235. Halliwell B, Gutteridge JMC. Free Radicals in Biology and Medicine. Oxford, UK: Biosciences, 2007, p. 851.
    Google Scholar
  • 236. Halliwell B, Rafter J, Jenner A. Health promotion by flavonoids, tocopherols, tocotrienols, and other phenols: direct or indirect effects? Antioxidant or not? Am J Clin Nutr 81: 268S–276S, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 237. Halliwell B, Zhao K, Whiteman M. The gastrointestinal tract: a major site of antioxidant action? Free Radic Res 33: 819–830, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 238. Hamanishi T, Furuta H, Kato H, Doi A, Tamai M, Shimomura H, Sakagashira S, Nishi M, Sasaki H, Sanke T, Nanjo K. Functional variants in the glutathione peroxidase-1 (GPx-1) gene are associated with increased intima-media thickness of carotid arteries and risk of macrovascular diseases in Japanese type 2 diabetic patients. Diabetes 53: 2455–2460, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 239. Han ES, Muller FL, Perez VI, Qi W, Liang H, Xi L, Fu C, Doyle E, Hickey M, Cornell J, Epstein CJ, Roberts LJ, Van Remmen H, Richardson A. The in vivo gene expression signature of oxidative stress. Physiol Genomics 34: 112–126, 2008.
    Link | ISI | Google Scholar
  • 240. Hanschmann EM, Lonn ME, Schutte LD, Funke M, Godoy JR, Eitner S, Hudemann C, Lillig CH. Both thioredoxin 2 and glutaredoxin 2 contribute to the reduction of the mitochondrial 2-Cys peroxiredoxin Prx3. J Biol Chem 285: 40699–40705, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 241. Hansen R, Saebo M, Skjelbred CF, Nexo BA, Hagen PC, Bock G, Bowitz Lothe IM, Johnson E, Aase S, Hansteen IL, Vogel U, Kure EH. GPX Pro198Leu and OGG1 Ser326Cys polymorphisms and risk of development of colorectal adenomas and colorectal cancer. Cancer Lett 229: 85–91, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 242. Haque ME, Asanuma M, Higashi Y, Miyazaki I, Tanaka K, Ogawa N. Overexpression of Cu-Zn superoxide dismutase protects neuroblastoma cells against dopamine cytotoxicity accompanied by increase in their glutathione level. Neurosci Res 47: 31–37, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 243. Haque R, Chun E, Howell JC, Sengupta T, Chen D, Kim H. MicroRNA-30b-mediated regulation of catalase expression in human ARPE-19 cells. PLoS One 7: e42542, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 244. Harmon JS, Bogdani M, Parazzoli SD, Mak SS, Oseid EA, Berghmans M, Leboeuf RC, Robertson RP. β-Cell-specific overexpression of glutathione peroxidase preserves intranuclear MafA and reverses diabetes in db/db mice. Endocrinology 150: 4855–4862, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 245. Harris IS, Treloar AE, Inoue S, Sasaki M, Gorrini C, Lee KC, Yung KY, Brenner D, Knobbe-Thomsen CB, Cox MA, Elia A, Berger T, Cescon DW, Adeoye A, Brustle A, Molyneux SD, Mason JM, Li WY, Yamamoto K, Wakeham A, Berman HK, Khokha R, Done SJ, Kavanagh TJ, Lam CW, Mak TW. Glutathione and thioredoxin antioxidant pathways synergize to drive cancer initiation and progression. Cancer Cell 27: 211–222, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 246. Harrison-Findik DD, Klein E, Crist C, Evans J, Timchenko N, Gollan J. Iron-mediated regulation of liver hepcidin expression in rats and mice is abolished by alcohol. Hepatology 46: 1979–1985, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 247. Hashizume K, Hirasawa M, Imamura Y, Noda S, Shimizu T, Shinoda K, Kurihara T, Noda K, Ozawa Y, Ishida S, Miyake Y, Shirasawa T, Tsubota K. Retinal dysfunction and progressive retinal cell death in SOD1-deficient mice. Am J Pathol 172: 1325–1331, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 248. Hassett P, Curley GF, Contreras M, Masterson C, Higgins BD, O'Brien T, Devaney J, O'Toole D, Laffey JG. Overexpression of pulmonary extracellular superoxide dismutase attenuates endotoxin-induced acute lung injury. Intensive Care Med 37: 1680–1687, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 249. Hatfield DL, Gladyshev VN. The Outcome of Selenium and Vitamin E Cancer Prevention Trial (SELECT) reveals the need for better understanding of selenium biology. Mol Interv 9: 18–21, 2009.
    Crossref | PubMed | Google Scholar
  • 250. Hatfield DL, Tsuji PA, Carlson BA, Gladyshev VN. Selenium and selenocysteine: roles in cancer, health, and development. Trends Biochem Sci 39: 112–120, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 251. Hatfield DL, Yoo MH, Carlson BA, Gladyshev VN. Selenoproteins that function in cancer prevention and promotion. Biochim Biophys Acta 1790: 1541–1545, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 252. Hayes JD, McMahon M. NRF2 and KEAP1 mutations: permanent activation of an adaptive response in cancer. Trends Biochem Sci 34: 176–188, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 253. Hebert-Schuster M, Fabre EE, Nivet-Antoine V. Catalase polymorphisms and metabolic diseases. Curr Opin Clin Nutr Metab Care 15: 397–402, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 254. Hellfritsch J, Kirsch J, Schneider M, Fluege T, Wortmann M, Frijhoff J, Dagnell M, Fey T, Esposito I, Kolle P, Pogoda K, Angeli JP, Ingold I, Kuhlencordt P, Ostman A, Pohl U, Conrad M, Beck H. Knockout of mitochondrial thioredoxin reductase stabilizes prolyl hydroxylase 2 and inhibits tumor growth and tumor-derived angiogenesis. Antioxid Redox Signal 22: 938–950, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 255. Henderson CJ, Wolf CR, Kitteringham N, Powell H, Otto D, Park BK. Increased resistance to acetaminophen hepatotoxicity in mice lacking glutathione S-transferase Pi. Proc Natl Acad Sci USA 97: 12741–12745, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 256. Herault O, Hope KJ, Deneault E, Mayotte N, Chagraoui J, Wilhelm BT, Cellot S, Sauvageau M, Andrade-Navarro MA, Hebert J, Sauvageau G. A role for GPx3 in activity of normal and leukemia stem cells. J Exp Med 209: 895–901, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 257. Hill KE, Motley AK, Winfrey VP, Burk RF. Selenoprotein P is the major selenium transport protein in mouse milk. PLoS One 9: e103486, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 258. Hill KE, Zhou J, Austin LM, Motley AK, Ham AJ, Olson GE, Atkins JF, Gesteland RF, Burk RF. The selenium-rich C-terminal domain of mouse selenoprotein P is necessary for the supply of selenium to brain and testis but not for the maintenance of whole body selenium. J Biol Chem 282: 10972–10980, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 259. Hill KE, Zhou J, McMahan WJ, Motley AK, Burk RF. Neurological dysfunction occurs in mice with targeted deletion of the selenoprotein P gene. J Nutr 134: 157–161, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 260. Hirata H, Cadet JL. Kainate-induced hippocampal DNA damage is attenuated in superoxide dismutase transgenic mice. Mol Brain Res 48: 145–148, 1997.
    Crossref | PubMed | Google Scholar
  • 261. Hiroi S, Harada H, Nishi H, Satoh M, Nagai R, Kimura A. Polymorphisms in the SOD2 and HLA-DRB1 genes are associated with nonfamilial idiopathic dilated cardiomyopathy in Japanese. Biochem Biophys Res Commun 261: 332–339, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 262. Hirono A, Sasaya-Hamada F, Kanno H, Fujii H, Yoshida T, Miwa S. A novel human catalase mutation (358 T–>del) causing Japanese-type acatalasemia. Blood Cells Mol Dis 21: 232–234, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 263. Hirose K, Longo DL, Oppenheim JJ, Matsushima K. Overexpression of mitochondrial manganese superoxide dismutase promotes the survival of tumor cells exposed to interleukin-1, tumor necrosis factor, selected anticancer drugs, and ionizing radiation. FASEB J 7: 361–368, 1993.
    Crossref | PubMed | ISI | Google Scholar
  • 264. Ho YS, Gargano M, Cao J, Bronson RT, Heimler I, Hutz RJ. Reduced fertility in female mice lacking copper-zinc superoxide dismutase. J Biol Chem 273: 7765–7769, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 265. Ho YS, Magnenat JL, Bronson RT, Cao J, Gargano M, Sugawara M, Funk CD. Mice deficient in cellular glutathione peroxidase develop normally and show no increased sensitivity to hyperoxia. J Biol Chem 272: 16644–16651, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 266. Ho YS, Vincent R, Dey MS, Slot JW, Crapo JD. Transgenic models for the study of lung antioxidant defense: enhanced manganese-containing superoxide dismutase activity gives partial protection to B6C3 hybrid mice exposed to hyperoxia. Am J Respir Cell Mol Biol 18: 538–547, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 267. Ho YS, Xiong Y, Ho DS, Gao J, Chua BH, Pai H, Mieyal JJ. Targeted disruption of the glutaredoxin 1 gene does not sensitize adult mice to tissue injury induced by ischemia/reperfusion and hyperoxia. Free Radic Biol Med 43: 1299–1312, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 268. Ho YS, Xiong Y, Ma W, Spector A, Ho DS. Mice lacking catalase develop normally but show differential sensitivity to oxidant tissue injury. J Biol Chem 279: 32804–32812, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 269. Hodara R, Weiss D, Joseph G, Velasquez-Castano JC, Landazuri N, Han JW, Yoon YS, Taylor WR. Overexpression of catalase in myeloid cells causes impaired postischemic neovascularization. Arterioscler Thromb Vasc Biol 31: 2203–2209, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 270. Hoffman SM, Tully JE, Lahue KG, Anathy V, Nolin JD, Guala AS, van der Velden JL, Ho YS, Aliyeva M, Daphtary N, Lundblad LK, Irvin CG, Janssen-Heininger YM. Genetic ablation of glutaredoxin-1 causes enhanced resolution of airways hyperresponsiveness and mucus metaplasia in mice with allergic airways disease. Am J Physiol Lung Cell Mol Physiol 303: L528–L538, 2012.
    Link | ISI | Google Scholar
  • 271. Hofmann B, Hecht HJ, Flohe L. Peroxiredoxins. Biol Chem 383: 347–364, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 272. Hogan B, Beddington R, Constantini F, Lacy E. Manipulating the Mouse eEmbryo: A Laboratory Manual. New York: Cold Spring Harbor Laboratory Press, 1994.
    Google Scholar
  • 273. Hohmeier HE, Thigpen A, Tran VV, Davis R, Newgard CB. Stable expression of manganese superoxide dismutase (MnSOD) in insulinoma cells prevents IL-1beta-induced cytotoxicity and reduces nitric oxide production. J Clin Invest 101: 1811–1820, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 274. Holmgren A. Antioxidant function of thioredoxin and glutaredoxin systems. Antioxid Redox Signal 2: 811–820, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 275. Holmgren A. Hydrogen donor system for Escherichia coli ribonucleoside-diphosphate reductase dependent upon glutathione. Proc Natl Acad Sci USA 73: 2275–2279, 1976.
    Crossref | PubMed | ISI | Google Scholar
  • 276. Holmgren A. Thioredoxin. Annu Rev Biochem 54: 237–271, 1985.
    Crossref | PubMed | ISI | Google Scholar
  • 277. Holmgren A, Johansson C, Berndt C, Lonn ME, Hudemann C, Lillig CH. Thiol redox control via thioredoxin and glutaredoxin systems. Biochem Soc Trans 33: 1375–1377, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 278. Homma K, Fujisawa T, Tsuburaya N, Yamaguchi N, Kadowaki H, Takeda K, Nishitoh H, Matsuzawa A, Naguro I, Ichijo H. SOD1 as a molecular switch for initiating the homeostatic ER stress response under zinc deficiency. Mol Cell 52: 75–86, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 279. Hornberger TA, McLoughlin TJ, Leszczynski JK, Armstrong DD, Jameson RR, Bowen PE, Hwang ES, Hou H, Moustafa ME, Carlson BA, Hatfield DL, Diamond AM, Esser KA. Selenoprotein-deficient transgenic mice exhibit enhanced exercise-induced muscle growth. J Nutr 133: 3091–3097, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 280. Horstkotte J, Perisic T, Schneider M, Lange P, Schroeder M, Kiermayer C, Hinkel R, Ziegler T, Mandal PK, David R, Schulz S, Schmitt S, Widder J, Sinowatz F, Becker BF, Bauersachs J, Naebauer M, Franz WM, Jeremias I, Brielmeier M, Zischka H, Conrad M, Kupatt C. Mitochondrial thioredoxin reductase is essential for early postischemic myocardial protection. Circulation 124: 2892–2902, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 281. Hotta M, Tashiro F, Ikegami H, Niwa H, Ogihara T, Yodoi J, Miyazaki J. Pancreatic beta cell-specific expression of thioredoxin, an antioxidative and antiapoptotic protein, prevents autoimmune and streptozotocin-induced diabetes. J Exp Med 188: 1445–1451, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 282. Huang JQ, Li DL, Zhao H, Sun LH, Xia XJ, Wang KN, Luo X, Lei XG. The selenium deficiency disease exudative diathesis in chicks is associated with downregulation of seven common selenoprotein genes in liver and muscle. J Nutr 141: 1605–1610, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 283. Huang Q, Zhou HJ, Zhang H, Huang Y, Hinojosa-Kirschenbaum F, Fan P, Yao L, Belardinelli L, Tellides G, Giordano FJ, Budas GR, Min W. Thioredoxin-2 inhibits mitochondrial reactive oxygen species generation and apoptosis stress kinase-1 activity to maintain cardiac function. Circulation 131: 1082–1097, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 284. Huang TT, Carlson EJ, Gillespie AM, Shi Y, Epstein CJ. Ubiquitous overexpression of CuZn superoxide dismutase does not extend life span in mice. J Gerontol A Biol Sci Med Sci 55: B5–9, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 285. Huang TT, Yasunami M, Carlson EJ, Gillespie AM, Reaume AG, Hoffman EK, Chan PH, Scott RW, Epstein CJ. Superoxide-mediated cytotoxicity in superoxide dismutase-deficient fetal fibroblasts. Arch Biochem Biophys 344: 424–432, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 286. Huang TT, Zou Y, Corniola R. Oxidative stress and adult neurogenesis—effects of radiation and superoxide dismutase deficiency. Semin Cell Dev Biol 23: 738–744, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 287. Hudak BB, Zhang LY, Kleeberger SR. Inter-strain variation in susceptibility to hyperoxic injury of murine airways. Pharmacogenetics 3: 135–143, 1993.
    Crossref | PubMed | Google Scholar
  • 288. Hudson TS, Carlson BA, Hoeneroff MJ, Young HA, Sordillo L, Muller WJ, Hatfield DL, Green JE. Selenoproteins reduce susceptibility to DMBA-induced mammary carcinogenesis. Carcinogenesis 33: 1225–1230, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 289. Hwang I, Lee J, Huh JY, Park J, Lee HB, Ho YS, Ha H. Catalase deficiency accelerates diabetic renal injury through peroxisomal dysfunction. Diabetes 61: 728–738, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 290. Ichijo H, Nishida E, Irie K, ten Dijke P, Saitoh M, Moriguchi T, Takagi M, Matsumoto K, Miyazono K, Gotoh Y. Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science 275: 90–94, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 291. Ikegami T, Suzuki Y, Shimizu T, Isono K, Koseki H, Shirasawa T. Model mice for tissue-specific deletion of the manganese superoxide dismutase (MnSOD) gene. Biochem Biophys Res Commun 296: 729–736, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 292. Imai H, Hakkaku N, Iwamoto R, Suzuki J, Suzuki T, Tajima Y, Konishi K, Minami S, Ichinose S, Ishizaka K, Shioda S, Arata S, Nishimura M, Naito S, Nakagawa Y. Depletion of selenoprotein GPx4 in spermatocytes causes male infertility in mice. J Biol Chem 284: 32522–32532, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 293. Imai H, Hirao F, Sakamoto T, Sekine K, Mizukura Y, Saito M, Kitamoto T, Hayasaka M, Hanaoka K, Nakagawa Y. Early embryonic lethality caused by targeted disruption of the mouse PHGPx gene. Biochem Biophys Res Commun 305: 278–286, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 294. Imam SZ, Newport GD, Itzhak Y, Cadet JL, Islam F, Slikker W Jr, Ali SF. Peroxynitrite plays a role in methamphetamine-induced dopaminergic neurotoxicity: evidence from mice lacking neuronal nitric oxide synthase gene or overexpressing copper-zinc superoxide dismutase. J Neurochem 76: 745–749, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 295. Imamura Y, Noda S, Hashizume K, Shinoda K, Yamaguchi M, Uchiyama S, Shimizu T, Mizushima Y, Shirasawa T, Tsubota K. Drusen, choroidal neovascularization, and retinal pigment epithelium dysfunction in SOD1-deficient mice: a model of age-related macular degeneration. Proc Natl Acad Sci USA 103: 11282–11287, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 296. Irons R, Carlson BA, Hatfield DL, Davis CD. Both selenoproteins and low molecular weight selenocompounds reduce colon cancer risk in mice with genetically impaired selenoprotein expression. J Nutr 136: 1311–1317, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 297. Ischiropoulos H. Biological tyrosine nitration: a pathophysiological function of nitric oxide and reactive oxygen species. Arch Biochem Biophys 356: 1–11, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 298. Ischiropoulos H, Zhu L, Chen J, Tsai M, Martin JC, Smith CD, Beckman JS. Peroxynitrite-mediated tyrosine nitration catalyzed by superoxide dismutase. Arch Biochem Biophys 298: 431–437, 1992.
    Crossref | PubMed | ISI | Google Scholar
  • 299. Iuchi Y, Okada F, Onuma K, Onoda T, Asao H, Kobayashi M, Fujii J. Elevated oxidative stress in erythrocytes due to a SOD1 deficiency causes anaemia and triggers autoantibody production. Biochem J 402: 219–227, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 300. Iuchi Y, Okada F, Tsunoda S, Kibe N, Shirasawa N, Ikawa M, Okabe M, Ikeda Y, Fujii J. Peroxiredoxin 4 knockout results in elevated spermatogenic cell death via oxidative stress. Biochem J 419: 149–158, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 301. Iuchi Y, Roy D, Okada F, Kibe N, Tsunoda S, Suzuki S, Takahashi M, Yokoyama H, Yoshitake J, Kondo S, Fujii J. Spontaneous skin damage and delayed wound healing in SOD1-deficient mice. Mol Cell Biochem 341: 181–194, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 302. Iverson SV, Eriksson S, Xu J, Prigge JR, Talago EA, Meade TA, Meade ES, Capecchi MR, Arner ES, Schmidt EE. A Txnrd1-dependent metabolic switch alters hepatic lipogenesis, glycogen storage, and detoxification. Free Radic Biol Med 63: 369–380, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 303. Jaarsma D, Haasdijk ED, Grashorn JA, Hawkins R, van Duijn W, Verspaget HW, London J, Holstege JC. Human Cu/Zn superoxide dismutase (SOD1) overexpression in mice causes mitochondrial vacuolization, axonal degeneration, and premature motoneuron death and accelerates motoneuron disease in mice expressing a familial amyotrophic lateral sclerosis mutant SOD1. Neurobiol Dis 7: 623–643, 2000.
    PubMed | ISI | Google Scholar
  • 304. Jackson RM, Helton ES, Viera L, Ohman T. Survival, lung injury, and lung protein nitration in heterozygous MnSOD knockout mice in hyperoxia. Exp Lung Res 25: 631–636, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 305. Jakupoglu C, Przemeck GK, Schneider M, Moreno SG, Mayr N, Hatzopoulos AK, de Angelis MH, Wurst W, Bornkamm GW, Brielmeier M, Conrad M. Cytoplasmic thioredoxin reductase is essential for embryogenesis but dispensable for cardiac development. Mol Cell Biol 25: 1980–1988, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 306. James LP, Mayeux PR, Hinson JA. Acetaminophen-induced hepatotoxicity. Drug Metab Dispos 31: 1499–1506, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 307. Jang HH, Lee KO, Chi YH, Jung BG, Park SK, Park JH, Lee JR, Lee SS, Moon JC, Yun JW, Choi YO, Kim WY, Kang JS, Cheong GW, Yun DJ, Rhee SG, Cho MJ, Lee SY. Two enzymes in one; two yeast peroxiredoxins display oxidative stress-dependent switching from a peroxidase to a molecular chaperone function. Cell 117: 625–635, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 308. Jang YC, Perez VI, Song W, Lustgarten MS, Salmon AB, Mele J, Qi W, Liu Y, Liang H, Chaudhuri A, Ikeno Y, Epstein CJ, Van Remmen H, Richardson A. Overexpression of Mn superoxide dismutase does not increase life span in mice. J Gerontol A Biol Sci Med Sci 64: 1114–1125, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 309. Jiang D, Akopian G, Ho YS, Walsh JP, Andersen JK. Chronic brain oxidation in a glutathione peroxidase knockout mouse model results in increased resistance to induced epileptic seizures. Exp Neurol 164: 257–268, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 310. Jin R, Gao Y, Zhang S, Teng F, Xu X, Aili A, Wang Y, Sun X, Pang X, Ge Q, Zhang Y. Trx1/TrxR1 system regulates post-selected DP thymocytes survival by modulating ASK1-JNK/p38 MAPK activities. Immunol Cell Biol. In press.
    ISI | Google Scholar
  • 311. Jin RC, Mahoney CE, Coleman Anderson L, Ottaviano F, Croce K, Leopold JA, Zhang YY, Tang SS, Handy DE, Loscalzo J. Glutathione peroxidase-3 deficiency promotes platelet-dependent thrombosis in vivo. Circulation 123: 1963–1973, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 312. Kamezaki F, Tasaki H, Yamashita K, Tsutsui M, Koide S, Nakata S, Tanimoto A, Okazaki M, Sasaguri Y, Adachi T, Otsuji Y. Gene transfer of extracellular superoxide dismutase ameliorates pulmonary hypertension in rats. Am J Respir Crit Care Med 177: 219–226, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 313. Kamii H, Kato I, Kinouchi H, Chan PH, Epstein CJ, Akabane A, Okamoto H, Yoshimoto T. Amelioration of vasospasm after subarachnoid hemorrhage in transgenic mice overexpressing CuZn-superoxide dismutase. Stroke 30: 867–872, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 314. Kamimoto Y, Sugiyama T, Kihira T, Zhang L, Murabayashi N, Umekawa T, Nagao K, Ma N, Toyoda N, Yodoi J, Sagawa N. Transgenic mice overproducing human thioredoxin-1, an antioxidative and anti-apoptotic protein, prevents diabetic embryopathy. Diabetologia 53: 2046–2055, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 315. Kamsler A, Avital A, Greenberger V, Segal M. Aged SOD overexpressing mice exhibit enhanced spatial memory while lacking hippocampal neurogenesis. Antioxid Redox Signal 9: 181–189, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 316. Kamsler A, Segal M. Paradoxical actions of hydrogen peroxide on long-term potentiation in transgenic superoxide dismutase-1 mice. J Neurosci 23: 10359–10367, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 317. Kandadi MR, Yu X, Frankel AE, Ren J. Cardiac-specific catalase overexpression rescues anthrax lethal toxin-induced cardiac contractile dysfunction: role of oxidative stress and autophagy. BMC Med 10: 134, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 318. Kang SK, Rabbani ZN, Folz RJ, Golson ML, Huang H, Yu D, Samulski TS, Dewhirst MW, Anscher MS, Vujaskovic Z. Overexpression of extracellular superoxide dismutase protects mice from radiation-induced lung injury. Int J Radiat Oncol Biol Phys 57: 1056–1066, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 319. Kang YJ, Chen Y, Epstein PN. Suppression of doxorubicin cardiotoxicity by overexpression of catalase in the heart of transgenic mice. J Biol Chem 271: 12610–12616, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 320. Kang YJ, Sun X, Chen Y, Zhou Z. Inhibition of doxorubicin chronic toxicity in catalase-overexpressing transgenic mouse hearts. Chem Res Toxicol 15: 1–6, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 321. Kannan S, Muthusamy VR, Whitehead KJ, Wang L, Gomes AV, Litwin SE, Kensler TW, Abel ED, Hoidal JR, Rajasekaran NS. Nrf2 deficiency prevents reductive stress-induced hypertrophic cardiomyopathy. Cardiovasc Res 100: 63–73, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 322. Kasaikina MV, Hatfield DL, Gladyshev VN. Understanding selenoprotein function and regulation through the use of rodent models. Biochim Biophys Acta 1823: 1633–1642, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 323. Kasaikina MV, Lobanov AV, Malinouski MY, Lee BC, Seravalli J, Fomenko DE, Turanov AA, Finney L, Vogt S, Park TJ, Miller RA, Hatfield DL, Gladyshev VN. Reduced utilization of selenium by naked mole rats due to a specific defect in GPx1 expression. J Biol Chem 286: 17005–17014, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 324. Kasaikina MV, Turanov AA, Avanesov A, Schweizer U, Seeher S, Bronson RT, Novoselov SN, Carlson BA, Hatfield DL, Gladyshev VN. Contrasting roles of dietary selenium and selenoproteins in chemically induced hepatocarcinogenesis. Carcinogenesis 34: 1089–1095, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 325. Kawai H, Ota T, Suzuki F, Tatsuka M. Molecular cloning of mouse thioredoxin reductases. Gene 242: 321–330, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 326. Kawakami T, Puri N, Sodhi K, Bellner L, Takahashi T, Morita K, Rezzani R, Oury TD, Abraham NG. Reciprocal effects of oxidative stress on heme oxygenase expression and activity contributes to reno-vascular abnormalities in EC-SOD knockout mice. Int J Hypertens 2012: 740203, 2012.
    Crossref | PubMed | Google Scholar
  • 327. Kawatani Y, Suzuki T, Shimizu R, Kelly VP, Yamamoto M. Nrf2 and selenoproteins are essential for maintaining oxidative homeostasis in erythrocytes and protecting against hemolytic anemia. Blood 117: 986–996, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 328. Kensler TW, Egner PA, Agyeman AS, Visvanathan K, Groopman JD, Chen JG, Chen TY, Fahey JW, Talalay P. Keap1-nrf2 signaling: a target for cancer prevention by sulforaphane. Top Curr Chem 329: 163–177, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 329. Kessova IG, Ho YS, Thung S, Cederbaum AI. Alcohol-induced liver injury in mice lacking Cu, Zn-superoxide dismutase. Hepatology 38: 1136–1145, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 330. Kiermayer C, Michalke B, Schmidt J, Brielmeier M. Effect of selenium on thioredoxin reductase activity in Txnrd1 or Txnrd2 hemizygous mice. Biol Chem 388: 1091–1097, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 331. Kim HS, Ullevig SL, Zamora D, Lee CF, Asmis R. Redox regulation of MAPK phosphatase 1 controls monocyte migration and macrophage recruitment. Proc Natl Acad Sci USA 109: E2803–2812, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 332. Kim SH, Kim MO, Gao P, Youm CA, Park HR, Lee TS, Kim KS, Suh JG, Lee HT, Park BJ, Ryoo ZY, Lee TH. Overexpression of extracellular superoxide dismutase (EC-SOD) in mouse skin plays a protective role in DMBA/TPA-induced tumor formation. Oncol Res 15: 333–341, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 333. Kim SJ, Lee JW, Jung YS, Kwon do Y, Park HK, Ryu CS, Kim SK, Oh GT, Kim YC. Ethanol-induced liver injury and changes in sulfur amino acid metabolomics in glutathione peroxidase and catalase double knockout mice. J Hepatol 50: 1184–1191, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 334. Kimura N, Tsunoda S, Iuchi Y, Abe H, Totsukawa K, Fujii J. Intrinsic oxidative stress causes either 2-cell arrest or cell death depending on developmental stage of the embryos from SOD1-deficient mice. Mol Hum Reprod 16: 441–451, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 335. Kinugawa S, Wang Z, Kaminski PM, Wolin MS, Edwards JG, Kaley G, Hintze TH. Limited exercise capacity in heterozygous manganese superoxide dismutase gene-knockout mice: roles of superoxide anion and nitric oxide. Circulation 111: 1480–1486, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 336. Kipp AP, Muller MF, Goken EM, Deubel S, Brigelius-Flohe R. The selenoproteins GPx2, TrxR2 and TrxR3 are regulated by Wnt signalling in the intestinal epithelium. Biochim Biophys Acta 1820: 1588–1596, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 337. Kirkman HN, Gaetani GF. Mammalian catalase: a venerable enzyme with new mysteries. Trends Biochem Sci 32: 44–50, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 338. Kisucka J, Chauhan AK, Patten IS, Yesilaltay A, Neumann C, Van Etten RA, Krieger M, Wagner DD. Peroxiredoxin1 prevents excessive endothelial activation and early atherosclerosis. Circ Res 103: 598–605, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 340. Klein EA, Thompson IM Jr, Tangen CM, Crowley JJ, Lucia MS, Goodman PJ, Minasian LM, Ford LG, Parnes HL, Gaziano JM, Karp DD, Lieber MM, Walther PJ, Klotz L, Parsons JK, Chin JL, Darke AK, Lippman SM, Goodman GE, Meyskens FL Jr, Baker LH. Vitamin E and the risk of prostate cancer: the Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 306: 1549–1556, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 341. Klivenyi P, Andreassen OA, Ferrante RJ, Dedeoglu A, Mueller G, Lancelot E, Bogdanov M, Andersen JK, Jiang D, Beal MF. Mice deficient in cellular glutathione peroxidase show increased vulnerability to malonate, 3-nitropropionic acid, and 1-methyl-4-phenyl-1,2,5,6-tetrahydropyridine. J Neurosci 20: 1–7, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 342. Klivenyi P, St Clair D, Wermer M, Yen HC, Oberley T, Yang L, Flint Beal M. Manganese superoxide dismutase overexpression attenuates MPTP toxicity. Neurobiol Dis 5: 253–258, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 343. Knight TR, Ho YS, Farhood A, Jaeschke H. Peroxynitrite is a critical mediator of acetaminophen hepatotoxicity in murine livers: protection by glutathione. J Pharmacol Exp Ther 303: 468–475, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 344. Kobayashi M, Sugiyama H, Wang DH, Toda N, Maeshima Y, Yamasaki Y, Masuoka N, Yamada M, Kira S, Makino H. Catalase deficiency renders remnant kidneys more susceptible to oxidant tissue injury and renal fibrosis in mice. Kidney Int 68: 1018–1031, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 345. Kofler J, Hurn PD, Traystman RJ. SOD1 overexpression and female sex exhibit region-specific neuroprotection after global cerebral ischemia due to cardiac arrest. J Cereb Blood Flow Metab 25: 1130–1137, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 346. Kojima T, Wakamatsu TH, Dogru M, Ogawa Y, Igarashi A, Ibrahim OM, Inaba T, Shimizu T, Noda S, Obata H, Nakamura S, Wakamatsu A, Shirasawa T, Shimazaki J, Negishi K, Tsubota K. Age-related dysfunction of the lacrimal gland and oxidative stress: evidence from the Cu,Zn-superoxide dismutase-1 (Sod1) knockout mice. Am J Pathol 180: 1879–1896, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 347. Kondo T, Sharp FR, Honkaniemi J, Mikawa S, Epstein CJ, Chan PH. DNA fragmentation and Prolonged expression of c-fos, c-jun, and hsp70 in kainic acid-induced neuronal cell death in transgenic mice overexpressing human CuZn-superoxide dismutase. J Cereb Blood Flow Metab 17: 241–256, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 348. Kostrominova TY. Advanced age-related denervation and fiber-type grouping in skeletal muscle of SOD1 knockout mice. Free Radic Biol Med 49: 1582–1593, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 349. Kostrominova TY, Pasyk KA, Van Remmen H, Richardson AG, Faulkner JA. Adaptive changes in structure of skeletal muscles from adult Sod1 homozygous knockout mice. Cell Tissue Res 327: 595–605, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 350. Kotulska K, LePecheur M, Marcol W, Lewin-Kowalik J, Larysz-Brysz M, Paly E, Matuszek I, London J. Overexpression of copper/zinc-superoxide dismutase in transgenic mice markedly impairs regeneration and increases development of neuropathic pain after sciatic nerve injury. J Neurosci Res 84: 1091–1097, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 351. Kowluru RA, Kowluru V, Xiong Y, Ho YS. Overexpression of mitochondrial superoxide dismutase in mice protects the retina from diabetes-induced oxidative stress. Free Radic Biol Med 41: 1191–1196, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 352. Krehl S, Loewinger M, Florian S, Kipp AP, Banning A, Wessjohann LA, Brauer MN, Iori R, Esworthy RS, Chu FF, Brigelius-Flohe R. Glutathione peroxidase-2 and selenium decreased inflammation and tumors in a mouse model of inflammation-associated carcinogenesis whereas sulforaphane effects differed with selenium supply. Carcinogenesis 33: 620–628, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 353. Kryukov GV, Castellano S, Novoselov SV, Lobanov AV, Zehtab O, Guigo R, Gladyshev VN. Characterization of mammalian selenoproteomes. Science 300: 1439–1443, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 354. Kubisch HM, Wang J, Bray TM, Phillips JP. Targeted overexpression of Cu/Zn superoxide dismutase protects pancreatic beta-cells against oxidative stress. Diabetes 46: 1563–1566, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 355. Kubo E, Fatma N, Akagi Y, Beier DR, Singh SP, Singh DP. TAT-mediated PRDX6 protein transduction protects against eye lens epithelial cell death and delays lens opacity. Am J Physiol Cell Physiol 294: C842–C855, 2008.
    Link | ISI | Google Scholar
  • 356. Kumaraswamy E, Carlson BA, Morgan F, Miyoshi K, Robinson GW, Su D, Wang S, Southon E, Tessarollo L, Lee BJ, Gladyshev VN, Hennighausen L, Hatfield DL. Selective removal of the selenocysteine tRNA [Ser]Sec gene (Trsp) in mouse mammary epithelium. Mol Cell Biol 23: 1477–1488, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 357. Kunishige M, Hill KA, Riemer AM, Farwell KD, Halangoda A, Heinmoller E, Moore SR, Turner DM, Sommer SS. Mutation frequency is reduced in the cerebellum of Big Blue mice overexpressing a human wild type SOD1 gene. Mutat Res 473: 139–149, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 358. Kuo MD, Bright IJ, Wang DS, Ghafouri P, Yuksel E, Hilfiker PR, Miniati DN, Dake MD. Local resistance to oxidative stress by overexpression of copper-zinc superoxide dismutase limits neointimal formation after angioplasty. J Endovasc Ther 11: 585–594, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 359. Kuriakose GC, Kurup MG. Evaluation of renoprotective effect of Aphanizomenon flos-aquae on cisplatin-induced renal dysfunction in rats. Ren Fail 30: 717–725, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 360. Kurokawa S, Berry MJ. Selenium. Role of the essential metalloid in health. Met Ions Life Sci 13: 499–534, 2013.
    Crossref | PubMed | Google Scholar
  • 361. Labunskyy VM, Hatfield DL, Gladyshev VN. Selenoproteins: molecular pathways and physiological roles. Physiol Rev 94: 739–777, 2014.
    Link | ISI | Google Scholar
  • 362. Labunskyy VM, Lee BC, Handy DE, Loscalzo J, Hatfield DL, Gladyshev VN. Both maximal expression of selenoproteins and selenoprotein deficiency can promote development of type 2 diabetes-like phenotype in mice. Antioxid Redox Signal 14: 2327–2336, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 363. Laclaustra M, Navas-Acien A, Stranges S, Ordovas JM, Guallar E. Serum selenium concentrations and diabetes in US adults: National Health and Nutrition Examination Survey (NHANES) 2003–2004. Environ Health Perspect 117: 1409–1413, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 364. Lafon-Cazal M, Culcasi M, Gaven F, Pietri S, Bockaert J. Nitric oxide, superoxide and peroxynitrite: putative mediators of NMDA-induced cell death in cerebellar granule cells. Neuropharmacology 32: 1259–1266, 1993.
    Crossref | PubMed | ISI | Google Scholar
  • 365. Lafon-Cazal M, Pietri S, Culcasi M, Bockaert J. NMDA-dependent superoxide production and neurotoxicity. Nature 364: 535–537, 1993.
    Crossref | PubMed | ISI | Google Scholar
  • 366. Lana-Elola E, Watson-Scales SD, Fisher EM, Tybulewicz VL. Down syndrome: searching for the genetic culprits. Dis Model Mech 4: 586–595, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 367. Larkin LM, Davis CS, Sims-Robinson C, Kostrominova TY, Remmen HV, Richardson A, Feldman EL, Brooks SV. Skeletal muscle weakness due to deficiency of CuZn-superoxide dismutase is associated with loss of functional innervation. Am J Physiol Regul Integr Comp Physiol 301: R1400–R1407, 2011.
    Link | ISI | Google Scholar
  • 368. Larosche I, Choumar A, Fromenty B, Letteron P, Abbey-Toby A, Van Remmen H, Epstein CJ, Richardson A, Feldmann G, Pessayre D, Mansouri A. Prolonged ethanol administration depletes mitochondrial DNA in MnSOD-overexpressing transgenic mice, but not in their wild type littermates. Toxicol Appl Pharmacol 234: 326–338, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 369. Larsen GL, White CW, Takeda K, Loader JE, Nguyen DD, Joetham A, Groner Y, Gelfand EW. Mice that overexpress Cu/Zn superoxide dismutase are resistant to allergen-induced changes in airway control. Am J Physiol Lung Cell Mol Physiol 279: L350–L359, 2000.
    Link | ISI | Google Scholar
  • 370. Lebovitz RM, Zhang H, Vogel H, Cartwright J Jr, Dionne L, Lu N, Huang S, Matzuk MM. Neurodegeneration, myocardial injury, and perinatal death in mitochondrial superoxide dismutase-deficient mice. Proc Natl Acad Sci USA 93: 9782–9787, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 371. Lee BC, Peterfi Z, Hoffmann FW, Moore RE, Kaya A, Avanesov A, Tarrago L, Zhou Y, Weerapana E, Fomenko DE, Hoffmann PR, Gladyshev VN. MsrB1 and MICALs regulate actin assembly and macrophage function via reversible stereoselective methionine oxidation. Mol Cell 51: 397–404, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 372. Lee DH, Esworthy RS, Chu C, Pfeifer GP, Chu FF. Mutation accumulation in the intestine and colon of mice deficient in two intracellular glutathione peroxidases. Cancer Res 66: 9845–9851, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 373. Lee S, Kim HJ. Prion-like mechanism in amyotrophic lateral sclerosis: are protein aggregates the key? Exp Neurobiol 24: 1–7, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 374. Lee S, Van Remmen H, Csete M. Sod2 overexpression preserves myoblast mitochondrial mass and function, but not muscle mass with aging. Aging Cell 8: 296–310, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 375. Lee TH, Kim SU, Yu SL, Kim SH, Park DS, Moon HB, Dho SH, Kwon KS, Kwon HJ, Han YH, Jeong S, Kang SW, Shin HS, Lee KK, Rhee SG, Yu DY. Peroxiredoxin II is essential for sustaining life span of erythrocytes in mice. Blood 101: 5033–5038, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 376. Lei XG, Cheng WH. New roles for an old selenoenzyme: evidence from glutathione peroxidase-1 null and overexpressing mice. J Nutr 135: 2295–2298, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 377. Lei XG, Cheng WH, McClung JP. Metabolic regulation and function of glutathione peroxidase-1. Annu Rev Nutr 27: 41–61, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 378. Lei XG, Vatamaniuk MZ. Two tales of antioxidant enzymes on beta cells and diabetes. Antioxid Redox Signal 14: 489–503, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 379. Lei XG, Zhu JH, McClung JP, Aregullin M, Roneker CA. Mice deficient in Cu,Zn-superoxide dismutase are resistant to acetaminophen toxicity. Biochem J 399: 455–461, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 380. Leitzmann C. Other biologically active substances in plant foods. In: Essential of Human Nutrition, edited by Mann J, Truswell S. Oxford: Oxford University Press, 2002, p. 259–269.
    Google Scholar
  • 381. Levin ED, Christopher NC, Crapo JD. Memory decline of aging reduced by extracellular superoxide dismutase overexpression. Behav Genet 35: 447–453, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 382. Levin ED, Christopher NC, Lateef S, Elamir BM, Patel M, Liang LP, Crapo JD. Extracellular superoxide dismutase overexpression protects against aging-induced cognitive impairment in mice. Behav Genet 32: 119–125, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 383. Levy R, Glozman S, Milman D, Seruty C, Hagay Z, Yavin E, Groner Y. Ischemic reperfusion brain injury in fetal transgenic mice with elevated levels of copper-zinc superoxide dismutase. J Perinat Med 30: 158–165, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 384. Lewis P, Stefanovic N, Pete J, Calkin AC, Giunti S, Thallas-Bonke V, Jandeleit-Dahm KA, Allen TJ, Kola I, Cooper ME, de Haan JB. Lack of the antioxidant enzyme glutathione peroxidase-1 accelerates atherosclerosis in diabetic apolipoprotein E-deficient mice. Circulation 115: 2178–2187, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 385. Li F, Calingasan NY, Yu F, Mauck WM, Toidze M, Almeida CG, Takahashi RH, Carlson GA, Flint Beal M, Lin MT, Gouras GK. Increased plaque burden in brains of APP mutant MnSOD heterozygous knockout mice. J Neurochem 89: 1308–1312, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 386. Li G, Chen Y, Saari JT, Kang YJ. Catalase-overexpressing transgenic mouse heart is resistant to ischemia-reperfusion injury. Am J Physiol Heart Circ Physiol 273: H1090–H1095, 1997.
    Link | ISI | Google Scholar
  • 387. Li JJ, Oberley LW. Overexpression of manganese-containing superoxide dismutase confers resistance to the cytotoxicity of tumor necrosis factor alpha and/or hyperthermia. Cancer Res 57: 1991–1998, 1997.
    PubMed | ISI | Google Scholar
  • 388. Li JJ, Oberley LW, St Clair DK, Ridnour LA, Oberley TD. Phenotypic changes induced in human breast cancer cells by overexpression of manganese-containing superoxide dismutase. Oncogene 10: 1989–2000, 1995.
    PubMed | ISI | Google Scholar
  • 389. Li L, Shoji W, Takano H, Nishimura N, Aoki Y, Takahashi R, Goto S, Kaifu T, Takai T, Obinata M. Increased susceptibility of MER5 (peroxiredoxin III) knockout mice to LPS-induced oxidative stress. Biochem Biophys Res Commun 355: 715–721, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 390. Li X, Chen H, Epstein PN. Metallothionein and catalase sensitize to diabetes in nonobese diabetic mice: reactive oxygen species may have a protective role in pancreatic beta-cells. Diabetes 55: 1592–1604, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 391. Li X, Weng H, Reece EA, Yang P. SOD1 overexpression in vivo blocks hyperglycemia-induced specific PKC isoforms: substrate activation and consequent lipid peroxidation in diabetic embryopathy. Am J Obstet Gynecol 205: e81–86, 2011.
    Crossref | ISI | Google Scholar
  • 392. Li Y, Huang TT, Carlson EJ, Melov S, Ursell PC, Olson JL, Noble LJ, Yoshimura MP, Berger C, Chan PH, Wallace DC, Epstein CJ. Dilated cardiomyopathy and neonatal lethality in mutant mice lacking manganese superoxide dismutase. Nat Genet 11: 376–381, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 393. Liang H, Van Remmen H, Frohlich V, Lechleiter J, Richardson A, Ran Q. Gpx4 protects mitochondrial ATP generation against oxidative damage. Biochem Biophys Res Commun 356: 893–898, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 394. Liang LP, Waldbaum S, Rowley S, Huang TT, Day BJ, Patel M. Mitochondrial oxidative stress and epilepsy in SOD2 deficient mice: attenuation by a lipophilic metalloporphyrin. Neurobiol Dis 45: 1068–1076, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 395. Lillig CH, Berndt C, Holmgren A. Glutaredoxin systems. Biochim Biophys Acta 1780: 1304–1317, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 396. Lillig CH, Holmgren A. Thioredoxin and related molecules–from biology to health and disease. Antioxid Redox Signal 9: 25–47, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 397. Lim CC, Bryan NS, Jain M, Garcia-Saura MF, Fernandez BO, Sawyer DB, Handy DE, Loscalzo J, Feelisch M, Liao R. Glutathione peroxidase deficiency exacerbates ischemia-reperfusion injury in male but not female myocardium: insights into antioxidant compensatory mechanisms. Am J Physiol Heart Circ Physiol 297: H2144–H2153, 2009.
    Link | ISI | Google Scholar
  • 398. Liochev SI, Fridovich I. CO2, not HCO3, facilitates oxidations by Cu,Zn superoxide dismutase plus H2O2. Proc Natl Acad Sci USA 101: 743–744, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 399. Liochev SI, Fridovich I. Mechanism of the peroxidase activity of Cu, Zn superoxide dismutase. Free Radic Biol Med 48: 1565–1569, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 400. Lippman SM, Klein EA, Goodman PJ, Lucia MS, Thompson IM, Ford LG, Parnes HL, Minasian LM, Gaziano JM, Hartline JA, Parsons JK, Bearden JD, 3rd Crawford ED, Goodman GE, Claudio J, Winquist E, Cook ED, Karp DD, Walther P, Lieber MM, Kristal AR, Darke AK, Arnold KB, Ganz PA, Santella RM, Albanes D, Taylor PR, Probstfield JL, Jagpal TJ, Crowley JJ, Meyskens FL Jr, Baker LH, Coltman CA Jr. Effect of selenium and vitamin E on risk of prostate cancer and other cancers: the Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 301: 39–51, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 401. Litvan J, Briva A, Wilson MS, Budinger GR, Sznajder JI, Ridge KM. Beta-adrenergic receptor stimulation and adenoviral overexpression of superoxide dismutase prevent the hypoxia-mediated decrease in Na,K-ATPase and alveolar fluid reabsorption. J Biol Chem 281: 19892–19898, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 402. Liu C, Wu J, Zou MH. Activation of AMP-activated protein kinase alleviates high-glucose-induced dysfunction of brain microvascular endothelial cell tight-junction dynamics. Free Radical Biol Med 53: 1213–1221, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 403. Liu J, Hinkhouse MM, Sun W, Weydert CJ, Ritchie JM, Oberley LW, Cullen JJ. Redox regulation of pancreatic cancer cell growth: role of glutathione peroxidase in the suppression of the malignant phenotype. Hum Gene Ther 15: 239–250, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 404. Lo Conte M, Carroll KS. The redox biochemistry of protein sulfenylation and sulfinylation. J Biol Chem 288: 26480–26488, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 405. Locy ML, Rogers LK, Prigge JR, Schmidt EE, Arner ES, Tipple TE. Thioredoxin reductase inhibition elicits Nrf2-mediated responses in Clara cells: implications for oxidant-induced lung injury. Antioxid Redox Signal 17: 1407–1416, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 406. Loh K, Deng H, Fukushima A, Cai X, Boivin B, Galic S, Bruce C, Shields BJ, Skiba B, Ooms LM, Stepto N, Wu B, Mitchell CA, Tonks NK, Watt MJ, Febbraio MA, Crack PJ, Andrikopoulos S, Tiganis T. Reactive oxygen species enhance insulin sensitivity. Cell Metab 10: 260–272, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 407. Lortz S, Gurgul-Convey E, Naujok O, Lenzen S. Overexpression of the antioxidant enzyme catalase does not interfere with the glucose responsiveness of insulin-secreting INS-1E cells and rat islets. Diabetologia 56: 774–782, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 408. Lortz S, Tiedge M. Sequential inactivation of reactive oxygen species by combined overexpression of SOD isoforms and catalase in insulin-producing cells. Free Radical Biol Med 34: 683–688, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 409. Low FM, Hampton MB, Peskin AV, Winterbourn CC. Peroxiredoxin 2 functions as a noncatalytic scavenger of low-level hydrogen peroxide in the erythrocyte. Blood 109: 2611–2617, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 410. Lu J, Chew EH, Holmgren A. Targeting thioredoxin reductase is a basis for cancer therapy by arsenic trioxide. Proc Natl Acad Sci USA 104: 12288–12293, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 411. Lu J, Holmgren A. The thioredoxin antioxidant system. Free Radical Biol Med 66: 75–87, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 412. Lu J, Holmgren A. Thioredoxin system in cell death progression. Antioxid Redox Signal 17: 1738–1747, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 413. Lu YP, Lou YR, Yen P, Newmark HL, Mirochnitchenko OI, Inouye M, Huang MT. Enhanced skin carcinogenesis in transgenic mice with high expression of glutathione peroxidase or both glutathione peroxidase and superoxide dismutase. Cancer Res 57: 1468–1474, 1997.
    PubMed | ISI | Google Scholar
  • 414. Lu Z, Xu X, Hu X, Zhu G, Zhang P, van Deel ED, French JP, Fassett JT, Oury TD, Bache RJ, Chen Y. Extracellular superoxide dismutase deficiency exacerbates pressure overload-induced left ventricular hypertrophy and dysfunction. Hypertension 51: 19–25, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 415. Luchman HA, Villemaire ML, Bismar TA, Carlson BA, Jirik FR. Prostate epithelium-specific deletion of the selenocysteine tRNA gene Trsp leads to early onset intraepithelial neoplasia. Am J Pathol 184: 871–877, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 416. Lupertz R, Chovolou Y, Kampkotter A, Watjen W, Kahl R. Catalase overexpression impairs TNF-alpha induced NF-kappaB activation and sensitizes MCF-7 cells against TNF-alpha. J Cell Biochem 103: 1497–1511, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 417. Lustgarten MS, Jang YC, Liu Y, Muller FL, Qi W, Steinhelper M, Brooks SV, Larkin L, Shimizu T, Shirasawa T, McManus LM, Bhattacharya A, Richardson A, Van Remmen H. Conditional knockout of Mn-SOD targeted to type IIB skeletal muscle fibers increases oxidative stress and is sufficient to alter aerobic exercise capacity. Am J Physiol Cell Physiol 297: C1520–C1532, 2009.
    Link | ISI | Google Scholar
  • 418. Lustgarten MS, Jang YC, Liu Y, Qi W, Qin Y, Dahia PL, Shi Y, Bhattacharya A, Muller FL, Shimizu T, Shirasawa T, Richardson A, Van Remmen H. MnSOD deficiency results in elevated oxidative stress and decreased mitochondrial function but does not lead to muscle atrophy during aging. Aging Cell 10: 493–505, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 419. Ma Q. Role of nrf2 in oxidative stress and toxicity. Annu Rev Pharmacol Toxicol 53: 401–426, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 420. Ma Q, He X. Molecular basis of electrophilic and oxidative defense: promises and perils of Nrf2. Pharmacol Rev 64: 1055–1081, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 421. MacGregor DG, Higgins MJ, Jones PA, Maxwell WL, Watson MW, Graham DI, Stone TW. Ascorbate attenuates the systemic kainate-induced neurotoxicity in the rat hippocampus. Brain Res 727: 133–144, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 422. Madamanchi NR, Moon SK, Hakim ZS, Clark S, Mehrizi A, Patterson C, Runge MS. Differential activation of mitogenic signaling pathways in aortic smooth muscle cells deficient in superoxide dismutase isoforms. Arterioscler Thromb Vasc Biol 25: 950–956, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 423. Mahmood DF, Abderrazak A, El Hadri K, Simmet T, Rouis M. The thioredoxin system as a therapeutic target in human health and disease. Antioxid Redox Signal 19: 1266–1303, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 424. Maiellaro-Rafferty K, Weiss D, Joseph G, Wan W, Gleason RL, Taylor WR. Catalase overexpression in aortic smooth muscle prevents pathological mechanical changes underlying abdominal aortic aneurysm formation. Am J Physiol Heart Circ Physiol 301: H355–H362, 2011.
    Link | ISI | Google Scholar
  • 425. Maier CM, Hsieh L, Crandall T, Narasimhan P, Chan PH. Evaluating therapeutic targets for reperfusion-related brain hemorrhage. Ann Neurol 59: 929–938, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 426. Mailloux RJ, Xuan JY, Beauchamp B, Jui L, Lou M, Harper ME. Glutaredoxin-2 is required to control proton leak through uncoupling protein-3. J Biol Chem 288: 8365–8379, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 427. Mailloux RJ, Xuan JY, McBride S, Maharsy W, Thorn S, Holterman CE, Kennedy CR, Rippstein P, deKemp R, da Silva J, Nemer M, Lou M, Harper ME. Glutaredoxin-2 is required to control oxidative phosphorylation in cardiac muscle by mediating deglutathionylation reactions. J Biol Chem 289: 14812–14828, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 428. Makino N, Mochizuki Y, Bannai S, Sugita Y. Kinetic studies on the removal of extracellular hydrogen peroxide by cultured fibroblasts. J Biol Chem 269: 1020–1025, 1994.
    PubMed | ISI | Google Scholar
  • 429. Malinouski M, Kehr S, Finney L, Vogt S, Carlson BA, Seravalli J, Jin R, Handy DE, Park TJ, Loscalzo J, Hatfield DL, Gladyshev VN. High-resolution imaging of selenium in kidneys: a localized selenium pool associated with glutathione peroxidase 3. Antioxid Redox Signal 16: 185–192, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 430. Mandal PK, Schneider M, Kolle P, Kuhlencordt P, Forster H, Beck H, Bornkamm GW, Conrad M. Loss of thioredoxin reductase 1 renders tumors highly susceptible to pharmacologic glutathione deprivation. Cancer Res 70: 9505–9514, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 431. Manevich Y, Townsend DM, Hutchens S, Tew KD. Diazeniumdiolate mediated nitrosative stress alters nitric oxide homeostasis through intracellular calcium and S-glutathionylation of nitric oxide synthetase. PLoS One 5: e14151, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 432. Manna SK, Zhang HJ, Yan T, Oberley LW, Aggarwal BB. Overexpression of manganese superoxide dismutase suppresses tumor necrosis factor-induced apoptosis and activation of nuclear transcription factor-kappaB and activated protein-1. J Biol Chem 273: 13245–13254, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 433. Mansouri A, Tarhuni A, Larosche I, Reyl-Desmars F, Demeilliers C, Degoul F, Nahon P, Sutton A, Moreau R, Fromenty B, Pessayre D. MnSOD overexpression prevents liver mitochondrial DNA depletion after an alcohol binge but worsens this effect after prolonged alcohol consumption in mice. Dig Dis 28: 756–775, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 434. Martin FM, Xu X, von Lohneysen K, Gilmartin TJ, Friedman JS. SOD2 deficient erythroid cells up-regulate transferrin receptor and down-regulate mitochondrial biogenesis and metabolism. PLoS One 6: e16894, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 435. Massaad CA, Washington TM, Pautler RG, Klann E. Overexpression of SOD-2 reduces hippocampal superoxide and prevents memory deficits in a mouse model of Alzheimer's disease. Proc Natl Acad Sci USA 106: 13576–13581, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 436. Massilamany C, Gangaplara A, Kim H, Stanford C, Rathnaiah G, Steffen D, Lee J, Reddy J. Copper-zinc superoxide dismutase-deficient mice show increased susceptibility to experimental autoimmune encephalomyelitis induced with myelin oligodendrocyte glycoprotein 35–55. J Neuroimmunol 256: 19–27, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 437. Matsui M, Oshima M, Oshima H, Takaku K, Maruyama T, Yodoi J, Taketo MM. Early embryonic lethality caused by targeted disruption of the mouse thioredoxin gene. Dev Biol 178: 179–185, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 439. Matsuo Y, Yodoi J. Extracellular thioredoxin: a therapeutic tool to combat inflammation. Cytokine Growth Factor Rev 24: 345–353, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 440. Matsushima S, Ide T, Yamato M, Matsusaka H, Hattori F, Ikeuchi M, Kubota T, Sunagawa K, Hasegawa Y, Kurihara T, Oikawa S, Kinugawa S, Tsutsui H. Overexpression of mitochondrial peroxiredoxin-3 prevents left ventricular remodeling and failure after myocardial infarction in mice. Circulation 113: 1779–1786, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 441. Matsushita M, Freigang S, Schneider C, Conrad M, Bornkamm GW, Kopf M. T cell lipid peroxidation induces ferroptosis and prevents immunity to infection. J Exp Med 212: 555–568, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 442. Matsuzaka Y, Okamoto K, Mabuchi T, Iizuka M, Ozawa A, Oka A, Tamiya G, Kulski JK, Inoko H. Identification and characterization of novel variants of the thioredoxin reductase 3 new transcript 1 TXNRD3NT1. Mamm Genome 16: 41–49, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 443. Matzuk MM, Dionne L, Guo Q, Kumar TR, Lebovitz RM. Ovarian function in superoxide dismutase 1 and 2 knockout mice. Endocrinology 139: 4008–4011, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 444. McCann JC, Ames BN. Adaptive dysfunction of selenoproteins from the perspective of the triage theory: why modest selenium deficiency may increase risk of diseases of aging. FASEB J 25: 1793–1814, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 445. McClung JP, Roneker CA, Mu W, Lisk DJ, Langlais P, Liu F, Lei XG. Development of insulin resistance and obesity in mice overexpressing cellular glutathione peroxidase. Proc Natl Acad Sci USA 101: 8852–8857, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 446. McFadden SL, Ding D, Burkard RF, Jiang H, Reaume AG, Flood DG, Salvi RJ. Cu/Zn SOD deficiency potentiates hearing loss and cochlear pathology in aged 129,CD-1 mice. J Comp Neurol 413: 101–112, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 447. McGirt MJ, Parra A, Sheng H, Higuchi Y, Oury TD, Laskowitz DT, Pearlstein RD, Warner DS. Attenuation of cerebral vasospasm after subarachnoid hemorrhage in mice overexpressing extracellular superoxide dismutase. Stroke 33: 2317–2323, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 448. Medinas DB, Toledo JC Jr, Cerchiaro G, do-Amaral AT, de-Rezende L, Malvezzi A, Augusto O. Peroxymonocarbonate and carbonate radical displace the hydroxyl-like oxidant in the Sod1 peroxidase activity under physiological conditions. Chem Res Toxicol 22: 639–648, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 449. Mendell JT, Olson EN. MicroRNAs in stress signaling and human disease. Cell 148: 1172–1187, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 450. Meng F, Yao D, Shi Y, Kabakoff J, Wu W, Reicher J, Ma Y, Moosmann B, Masliah E, Lipton SA, Gu Z. Oxidation of the cysteine-rich regions of parkin perturbs its E3 ligase activity and contributes to protein aggregation. Mol Neurodegener 6: 34, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 451. Meng TC, Fukada T, Tonks NK. Reversible oxidation and inactivation of protein tyrosine phosphatases in vivo. Mol Cell 9: 387–399, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 452. Milatovic D, Gupta RC, Dettbarn WD. Involvement of nitric oxide in kainic acid-induced excitotoxicity in rat brain. Brain Res 957: 330–337, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 453. Min D, Kim H, Park L, Kim TH, Hwang S, Kim MJ, Jang S, Park Y. Amelioration of diabetic neuropathy by TAT-mediated enhanced delivery of metallothionein and SOD. Endocrinology 153: 81–91, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 454. Miranda-Vizuete A, Damdimopoulos AE, Spyrou G. cDNA cloning, expression and chromosomal localization of the mouse mitochondrial thioredoxin reductase gene(1). Biochim Biophys Acta 1447: 113–118, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 455. Miranda-Vizuete A, Spyrou G. Genomic organization and identification of a novel alternative splicing variant of mouse mitochondrial thioredoxin reductase (TrxR2) gene. Mol Cell 13: 488–492, 2002.
    Google Scholar
  • 456. Mirochnitchenko O, Inouye M. Effect of overexpression of human Cu,Zn superoxide dismutase in transgenic mice on macrophage functions. J Immunol 156: 1578–1586, 1996.
    PubMed | ISI | Google Scholar
  • 457. Mirochnitchenko O, Palnitkar U, Philbert M, Inouye M. Thermosensitive phenotype of transgenic mice overproducing human glutathione peroxidases. Proc Natl Acad Sci USA 92: 8120–8124, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 458. Mirochnitchenko O, Weisbrot-Lefkowitz M, Reuhl K, Chen L, Yang C, Inouye M. Acetaminophen toxicity. Opposite effects of two forms of glutathione peroxidase. J Biol Chem 274: 10349–10355, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 459. Misawa H, Nakata K, Matsuura J, Moriwaki Y, Kawashima K, Shimizu T, Shirasawa T, Takahashi R. Conditional knockout of Mn superoxide dismutase in postnatal motor neurons reveals resistance to mitochondrial generated superoxide radicals. Neurobiol Dis 23: 169–177, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 460. Mitsuishi Y, Motohashi H, Yamamoto M. The Keap1-Nrf2 system in cancers: stress response and anabolic metabolism. Front Oncol 2: 200, 2012.
    Crossref | PubMed | Google Scholar
  • 461. Mo Y, Feinstein SI, Manevich Y, Zhang Q, Lu L, Ho YS, Fisher AB. 1-Cys peroxiredoxin knock-out mice express mRNA but not protein for a highly related intronless gene. FEBS Lett 555: 192–198, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 462. Mohr S, Hallak H, de Boitte A, Lapetina EG, Brune B. Nitric oxide-induced S-glutathionylation and inactivation of glyceraldehyde-3-phosphate dehydrogenase. J Biol Chem 274: 9427–9430, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 463. Moi P, Chan K, Asunis I, Cao A, Kan YW. Isolation of NF-E2-related factor 2 (Nrf2), a NF-E2-like basic leucine zipper transcriptional activator that binds to the tandem NF-E2/AP1 repeat of the beta-globin locus control region. Proc Natl Acad Sci USA 91: 9926–9930, 1994.
    Crossref | PubMed | ISI | Google Scholar
  • 464. Moon EJ, Giaccia A. Dual roles of NRF2 in tumor prevention and progression: possible implications in cancer treatment. Free Radic Biol Med 79: 292–299, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 465. Moon JC, Hah YS, Kim WY, Jung BG, Jang HH, Lee JR, Kim SY, Lee YM, Jeon MG, Kim CW, Cho MJ, Lee SY. Oxidative stress-dependent structural and functional switching of a human 2-Cys peroxiredoxin isotype II that enhances HeLa cell resistance to H2O2-induced cell death. J Biol Chem 280: 28775–28784, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 466. Morigasaki S, Shimada K, Ikner A, Yanagida M, Shiozaki K. Glycolytic enzyme GAPDH promotes peroxide stress signaling through multistep phosphorelay to a MAPK cascade. Mol Cell 30: 108–113, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 467. Morita-Fujimura Y, Fujimura M, Gasche Y, Copin JC, Chan PH. Overexpression of copper and zinc superoxide dismutase in transgenic mice prevents the induction and activation of matrix metalloproteinases after cold injury-induced brain trauma. J Cereb Blood Flow Metab 20: 130–138, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 468. Moskovitz J, Bar-Noy S, Williams WM, Requena J, Berlett BS, Stadtman ER. Methionine sulfoxide reductase (MsrA) is a regulator of antioxidant defense and lifespan in mammals. Proc Natl Acad Sci USA 98: 12920–12925, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 469. Motoori S, Majima HJ, Ebara M, Kato H, Hirai F, Kakinuma S, Yamaguchi C, Ozawa T, Nagano T, Tsujii H, Saisho H. Overexpression of mitochondrial manganese superoxide dismutase protects against radiation-induced cell death in the human hepatocellular carcinoma cell line HLE. Cancer Res 61: 5382–5388, 2001.
    PubMed | ISI | Google Scholar
  • 470. Moustafa ME, Carlson BA, Anver MR, Bobe G, Zhong N, Ward JM, Perella CM, Hoffmann VJ, Rogers K, Combs GF Jr, Schweizer U, Merlino G, Gladyshev VN, Hatfield DL. Selenium and selenoprotein deficiencies induce widespread pyogranuloma formation in mice, while high levels of dietary selenium decrease liver tumor size driven by TGFalpha. PLoS One 8: e57389, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 471. Moustafa ME, Carlson BA, El-Saadani MA, Kryukov GV, Sun QA, Harney JW, Hill KE, Combs GF, Feigenbaum L, Mansur DB, Burk RF, Berry MJ, Diamond AM, Lee BJ, Gladyshev VN, Hatfield DL. Selective inhibition of selenocysteine tRNA maturation and selenoprotein synthesis in transgenic mice expressing isopentenyladenosine-deficient selenocysteine tRNA. Mol Cell Biol 21: 3840–3852, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 472. Mulder DW. Clinical limits of amyotrophic lateral sclerosis. Adv Neurol 36: 15–22, 1982.
    PubMed | Google Scholar
  • 473. Muller MF, Florian S, Pommer S, Osterhoff M, Esworthy RS, Chu FF, Brigelius-Flohe R, Kipp AP. Deletion of glutathione peroxidase-2 inhibits azoxymethane-induced colon cancer development. PLoS One 8: e72055, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 474. Mulligan VK, Chakrabartty A. Protein misfolding in the late-onset neurodegenerative diseases: common themes and the unique case of amyotrophic lateral sclerosis. Proteins 81: 1285–1303, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 475. Muoio DM. TXNIP links redox circuitry to glucose control. Cell Metab 5: 412–414, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 476. Murakami K, Kondo T, Epstein CJ, Chan PH. Overexpression of CuZn-superoxide dismutase reduces hippocampal injury after global ischemia in transgenic mice. Stroke 28: 1797–1804, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 477. Murakami K, Murata N, Noda Y, Tahara S, Kaneko T, Kinoshita N, Hatsuta H, Murayama S, Barnham KJ, Irie K, Shirasawa T, Shimizu T. SOD1 (copper/zinc superoxide dismutase) deficiency drives amyloid beta protein oligomerization and memory loss in mouse model of Alzheimer disease. J Biol Chem 286: 44557–44568, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 478. Murdoch CE, Shuler M, Haeussler DJ, Kikuchi R, Bearelly P, Han J, Watanabe Y, Fuster JJ, Walsh K, Ho YS, Bachschmid MM, Cohen RA, Matsui R. Glutaredoxin-1 up-regulation induces soluble vascular endothelial growth factor receptor 1, attenuating post-ischemia limb revascularization. J Biol Chem 289: 8633–8644, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 479. Murphy SJ, Hughes AE, Patterson CC, Anderson LA, Watson RGP, Johnston BT, Comber H, McGuigan J, Reynolds JV, Murray LJ. A population-based association study of SNPs of GSTP1, MnSOD, GPX2 and Barrett's esophagus and esophageal adenocarcinoma. Carcinogenesis 28: 1323–1328, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 480. Mysore TB, Shinkel TA, Collins J, Salvaris EJ, Fisicaro N, Murray-Segal LJ, Johnson LE, Lepore DA, Walters SN, Stokes R, Chandra AP, O'Connell PJ, d'Apice AJ, Cowan PJ. Overexpression of glutathione peroxidase with two isoforms of superoxide dismutase protects mouse islets from oxidative injury and improves islet graft function. Diabetes 54: 2109–2116, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 481. Nahon P, Charnaux N, Friand V, Prost-Squarcioni C, Ziol M, Lievre N, Trinchet JC, Beaugrand M, Gattegno L, Pessayre D, Sutton A. The manganese superoxide dismutase Ala16Val dimorphism modulates iron accumulation in human hepatoma cells. Free Radic Biol Med 45: 1308–1317, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 482. Nakamura H, De Rosa S, Roederer M, Anderson MT, Dubs JG, Yodoi J, Holmgren A, Herzenberg LA. Elevation of plasma thioredoxin levels in HIV-infected individuals. Int Immunol 8: 603–611, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 483. Nakamura H, Hoshino Y, Okuyama H, Matsuo Y, Yodoi J. Thioredoxin 1 delivery as new therapeutics. Adv Drug Deliv Rev 61: 303–309, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 484. Neumann CA, Krause DS, Carman CV, Das S, Dubey DP, Abraham JL, Bronson RT, Fujiwara Y, Orkin SH, Van Etten RA. Essential role for the peroxiredoxin Prdx1 in erythrocyte antioxidant defence and tumour suppression. Nature 424: 561–565, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 485. Nguyen P, Awwad RT, Smart DD, Spitz DR, Gius D. Thioredoxin reductase as a novel molecular target for cancer therapy. Cancer Lett 236: 164–174, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 486. Nilakantan V, Li Y, Spear BT, Glauert HP. Increased liver-specific expression of catalase in transgenic mice. Ann NY Acad Sci 804: 542–553, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 487. Nishi T, Shimizu N, Hiramoto M, Sato I, Yamaguchi Y, Hasegawa M, Aizawa S, Tanaka H, Kataoka K, Watanabe H, Handa H. Spatial redox regulation of a critical cysteine residue of NF-kappa B in vivo. J Biol Chem 277: 44548–44556, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 488. No JH, Kim YB, Song YS. Targeting nrf2 signaling to combat chemoresistance. J Cancer Prev 19: 111–117, 2014.
    Crossref | PubMed | Google Scholar
  • 489. Noda Y, Ota K, Shirasawa T, Shimizu T. Copper/zinc superoxide dismutase insufficiency impairs progesterone secretion and fertility in female mice. Biol Reprod 86: 1–8, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 490. Nonn L, Williams RR, Erickson RP, Powis G. The absence of mitochondrial thioredoxin 2 causes massive apoptosis, exencephaly, and early embryonic lethality in homozygous mice. Mol Cell Biol 23: 916–922, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 491. Nordlund A, Oliveberg M. SOD1-associated ALS: a promising system for elucidating the origin of protein-misfolding disease. HFSP J 2: 354–364, 2008.
    Crossref | PubMed | Google Scholar
  • 492. Nozik-Grayck E, Suliman HB, Majka S, Albietz J, Van Rheen Z, Roush K, Stenmark KR. Lung EC-SOD overexpression attenuates hypoxic induction of Egr-1 and chronic hypoxic pulmonary vascular remodeling. Am J Physiol Lung Cell Mol Physiol 295: L422–L430, 2008.
    Link | ISI | Google Scholar
  • 493. Obal D, Dai S, Keith R, Dimova N, Kingery J, Zheng YT, Zweier J, Velayutham M, Prabhu SD, Li Q, Conklin D, Yang D, Bhatnagar A, Bolli R, Rokosh G. Cardiomyocyte-restricted overexpression of extracellular superoxide dismutase increases nitric oxide bioavailability and reduces infarct size after ischemia/reperfusion. Basic Res Cardiol 107: 305, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 494. Oberley TD, Coursin DB, Cihla HP, Oberley LW, el-Sayyad N, Ho YS. Immunolocalization of manganese superoxide dismutase in normal and transgenic mice expressing the human enzyme. Histochem J 25: 267–279, 1993.
    Crossref | PubMed | Google Scholar
  • 495. Ogata M, Wang DH, Ogino K. Mammalian acatalasemia: the perspectives of bioinformatics and genetic toxicology. Acta Med Okayama 62: 345–361, 2008.
    PubMed | ISI | Google Scholar
  • 496. Ogura T, Tong KI, Mio K, Maruyama Y, Kurokawa H, Sato C, Yamamoto M. Keap1 is a forked-stem dimer structure with two large spheres enclosing the intervening, double glycine repeat, and C-terminal domains. Proc Natl Acad Sci USA 107: 2842–2847, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 497. Oh SS, Sullivan KA, Wilkinson JE, Backus C, Hayes JM, Sakowski SA, Feldman EL. Neurodegeneration and early lethality in superoxide dismutase 2-deficient mice: a comprehensive analysis of the central and peripheral nervous systems. Neuroscience 212: 201–213, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 498. Ohashi M, Runge MS, Faraci FM, Heistad DD. MnSOD deficiency increases endothelial dysfunction in ApoE-deficient mice. Arterioscler Thromb Vasc Biol 26: 2331–2336, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 499. Oien DB, Moskovitz J. Selenium and the methionine sulfoxide reductase system. Molecules 14: 2337–2344, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 500. Olofsson EM, Marklund SL, Behndig A. Enhanced age-related cataract in copper-zinc superoxide dismutase null mice. Clin Experiment Ophthalmol 40: 813–820, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 501. Olofsson EM, Marklund SL, Behndig A. Enhanced diabetes-induced cataract in copper-zinc superoxide dismutase-null mice. Invest Ophthalmol Vis Sci 50: 2913–2918, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 502. Olofsson EM, Marklund SL, Behndig A. Glucose-induced cataract in CuZn-SOD null lenses: an effect of nitric oxide? Free Radic Biol Med 42: 1098–1105, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 503. Olofsson EM, Marklund SL, Karlsson K, Brannstrom T, Behndig A. In vitro glucose-induced cataract in copper-zinc superoxide dismutase null mice. Exp Eye Res 81: 639–646, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 504. Olsen RH, Johnson LA, Zuloaga DG, Limoli CL, Raber J. Enhanced hippocampus-dependent memory and reduced anxiety in mice over-expressing human catalase in mitochondria. J Neurochem 125: 303–313, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 505. Olson GE, Whitin JC, Hill KE, Winfrey VP, Motley AK, Austin LM, Deal J, Cohen HJ, Burk RF. Extracellular glutathione peroxidase (Gpx3) binds specifically to basement membranes of mouse renal cortex tubule cells. Am J Physiol Renal Physiol 298: F1244–F1253, 2010.
    Link | ISI | Google Scholar
  • 506. Opii WO, Joshi G, Head E, Milgram NW, Muggenburg BA, Klein JB, Pierce WM, Cotman CW, Butterfield DA. Proteomic identification of brain proteins in the canine model of human aging following a long-term treatment with antioxidants and a program of behavioral enrichment: relevance to Alzheimer's disease. Neurobiol Aging 29: 51–70, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 507. Osborne SA, Tonissen KF. Genomic organisation and alternative splicing of mouse and human thioredoxin reductase 1 genes. BMC Genomics 2: 10, 2001.
    Crossref | PubMed | Google Scholar
  • 508. Oshikawa J, Urao N, Kim HW, Kaplan N, Razvi M, McKinney R, Poole LB, Fukai T, Ushio-Fukai M. Extracellular SOD-derived H2O2 promotes VEGF signaling in caveolae/lipid rafts and post-ischemic angiogenesis in mice. PLoS One 5: e10189, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 509. Oury TD, Ho YS, Piantadosi CA, Crapo JD. Extracellular superoxide dismutase, nitric oxide, and central nervous system O2 toxicity. Proc Natl Acad Sci USA 89: 9715–9719, 1992.
    Crossref | PubMed | ISI | Google Scholar
  • 510. Oury TD, Piantadosi CA, Crapo JD. Cold-induced brain edema in mice. Involvement of extracellular superoxide dismutase and nitric oxide. J Biol Chem 268: 15394–15398, 1993.
    PubMed | ISI | Google Scholar
  • 511. Ozumi K, Tasaki H, Takatsu H, Nakata S, Morishita T, Koide S, Yamashita K, Tsutsui M, Okazaki M, Sasaguri Y, Adachi T, Nakashima Y. Extracellular superoxide dismutase overexpression reduces cuff-induced arterial neointimal formation. Atherosclerosis 181: 55–62, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 512. Pacher P, Beckman JS, Liaudet L. Nitric oxide and peroxynitrite in health and disease. Physiol Rev 87: 315–424, 2007.
    Link | ISI | Google Scholar
  • 513. Pahlavani MA, Mele JF, Richardson A. Effect of overexpression of human Cu/Zn-SOD on activation-induced lymphocyte proliferation and apoptosis. Free Radic Biol Med 30: 1319–1327, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 514. Palazzotti B, Pani G, Colavitti R, De Leo ME, Bedogni B, Borrello S, Galeotti T. Increased growth capacity of cervical-carcinoma cells over-expressing manganous superoxide dismutase. Int J Cancer 82: 145–150, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 515. Papaioannou V, Johnson R. Production of chimeras and genetically defined offspring from targeted ES cells.. In: Gene Targeting: A Practical Approach, edited by Joyner AL. New York: Oxford Univ. Press, 1993, p. 107–146.
    Google Scholar
  • 516. Parajuli N, Marine A, Simmons S, Saba H, Mitchell T, Shimizu T, Shirasawa T, Macmillan-Crow LA. Generation and characterization of a novel kidney-specific manganese superoxide dismutase knockout mouse. Free Radic Biol Med 51: 406–416, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 517. Park CK, Jung JH, Moon MJ, Kim YY, Kim JH, Park SH, Kim CY, Paek SH, Kim DG, Jung HW, Cho BK. Tissue expression of manganese superoxide dismutase is a candidate prognostic marker for glioblastoma. Oncology 77: 178–181, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 518. Park JG, Yoo JY, Jeong SJ, Choi JH, Lee MR, Lee MN, Hwa Lee J, Kim HC, Jo H, Yu DY, Kang SW, Rhee SG, Lee MH, Oh GT. Peroxiredoxin 2 deficiency exacerbates atherosclerosis in apolipoprotein E-deficient mice. Circ Res 109: 739–749, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 519. Park JW, Qi WN, Cai Y, Zelko I, Liu JQ, Chen LE, Urbaniak JR, Folz RJ. Skeletal muscle reperfusion injury is enhanced in extracellular superoxide dismutase knockout mouse. Am J Physiol Heart Circ Physiol 289: H181–H187, 2005.
    Link | ISI | Google Scholar
  • 520. Patterson AD, Carlson BA, Li F, Bonzo JA, Yoo MH, Krausz KW, Conrad M, Chen C, Gonzalez FJ, Hatfield DL. Disruption of thioredoxin reductase 1 protects mice from acute acetaminophen-induced hepatotoxicity through enhanced NRF2 activity. Chem Res Toxicol 26: 1088–1096, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 521. Pei ZM, Murata Y, Benning G, Thomine S, Klusener B, Allen GJ, Grill E, Schroeder JI. Calcium channels activated by hydrogen peroxide mediate abscisic acid signalling in guard cells. Nature 406: 731–734, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 522. Pekkari K, Goodarzi MT, Scheynius A, Holmgren A, Avila-Carino J. Truncated thioredoxin (Trx80) induces differentiation of human CD14+ monocytes into a novel cell type (TAMs) via activation of the MAP kinases p38, ERK, and JNK. Blood 105: 1598–1605, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 523. Pekkari K, Holmgren A. Truncated thioredoxin: physiological functions and mechanism. Antioxid Redox Signal 6: 53–61, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 524. Peled-Kamar M, Lotem J, Okon E, Sachs L, Groner Y. Thymic abnormalities and enhanced apoptosis of thymocytes and bone marrow cells in transgenic mice overexpressing Cu/Zn-superoxide dismutase: implications for Down syndrome. EMBO J 14: 4985–4993, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 525. Peled-Kamar M, Lotem J, Wirguin I, Weiner L, Hermalin A, Groner Y. Oxidative stress mediates impairment of muscle function in transgenic mice with elevated level of wild-type Cu/Zn superoxide dismutase. Proc Natl Acad Sci USA 94: 3883–3887, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 526. Pepper MP, Vatamaniuk MZ, Yan X, Roneker CA, Lei XG. Impacts of dietary selenium deficiency on metabolic phenotypes of diet-restricted GPX1-overexpressing mice. Antioxid Redox Signal 14: 383–390, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 527. Perez VI, Cortez LA, Lew CM, Rodriguez M, Webb CR, Van Remmen H, Chaudhuri A, Qi W, Lee S, Bokov A, Fok W, Jones D, Richardson A, Yodoi J, Zhang Y, Tominaga K, Hubbard GB, Ikeno Y. Thioredoxin 1 overexpression extends mainly the earlier part of life span in mice. J Gerontol A Biol Sci Med Sci 66: 1286–1299, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 528. Perveen S, Patel H, Arif A, Younis S, Codipilly CN, Ahmed M. Role of EC-SOD overexpression in preserving pulmonary angiogenesis inhibited by oxidative stress. PLoS One 7: e51945, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 529. Peter Y, Rotman G, Lotem J, Elson A, Shiloh Y, Groner Y. Elevated Cu/Zn-SOD exacerbates radiation sensitivity and hematopoietic abnormalities of Atm-deficient mice. EMBO J 20: 1538–1546, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 530. Pfeiffer S, Lass A, Schmidt K, Mayer B. Protein tyrosine nitration in mouse peritoneal macrophages activated in vitro and in vivo: evidence against an essential role of peroxynitrite. FASEB J 15: 2355–2364, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 531. Phelan SA, Wang X, Wallbrandt P, Forsman-Semb K, Paigen B. Overexpression of Prdx6 reduces H2O2 but does not prevent diet-induced atherosclerosis in the aortic root. Free Radic Biol Med 35: 1110–1120, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 532. Pineda JA, Aono M, Sheng H, Lynch J, Wellons JC, Laskowitz DT, Pearlstein RD, Bowler R, Crapo J, Warner DS. Extracellular superoxide dismutase overexpression improves behavioral outcome from closed head injury in the mouse. J Neurotrauma 18: 625–634, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 533. Pitts MW, Raman AV, Hashimoto AC, Todorovic C, Nichols RA, Berry MJ. Deletion of selenoprotein P results in impaired function of parvalbumin interneurons and alterations in fear learning and sensorimotor gating. Neuroscience 208: 58–68, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 534. Poyton RO, Castello PR, Ball KA, Woo DK, Pan N. Mitochondria and hypoxic signaling: a new view. Ann NY Acad Sci 1177: 48–56, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 535. Prigge JR, Eriksson S, Iverson SV, Meade TA, Capecchi MR, Arner ES, Schmidt EE. Hepatocyte DNA replication in growing liver requires either glutathione or a single allele of txnrd1. Free Radic Biol Med 52: 803–810, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 536. Przedborski S, Kostic V, Jackson-Lewis V, Naini AB, Simonetti S, Fahn S, Carlson E, Epstein CJ, Cadet JL. Transgenic mice with increased Cu/Zn-superoxide dismutase activity are resistant to N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced neurotoxicity. J Neurosci 12: 1658–1667, 1992.
    Crossref | PubMed | ISI | Google Scholar
  • 537. Qiao YL, Dawsey SM, Kamangar F, Fan JH, Abnet CC, Sun XD, Johnson LL, Gail MH, Dong ZW, Yu B, Mark SD, Taylor PR. Total and cancer mortality after supplementation with vitamins and minerals: follow-up of the Linxian General Population Nutrition Intervention Trial. J Natl Cancer Inst 101: 507–518, 2009.
    Crossref | PubMed | Google Scholar
  • 538. Qin F, Lennon-Edwards S, Lancel S, Biolo A, Siwik DA, Pimentel DR, Dorn GW, Kang YJ, Colucci WS. Cardiac-specific overexpression of catalase identifies hydrogen peroxide-dependent and -independent phases of myocardial remodeling and prevents the progression to overt heart failure in G(alpha)q-overexpressing transgenic mice. Circ Heart Fail 3: 306–313, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 539. Rabbani ZN, Anscher MS, Folz RJ, Archer E, Huang H, Chen L, Golson ML, Samulski TS, Dewhirst MW, Vujaskovic Z. Overexpression of extracellular superoxide dismutase reduces acute radiation induced lung toxicity. BMC Cancer 5: 59, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 540. Raber J, Villasana L, Rosenberg J, Zou Y, Huang TT, Fike JR. Irradiation enhances hippocampus-dependent cognition in mice deficient in extracellular superoxide dismutase. Hippocampus 21: 72–80, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 541. Radi R, Turrens JF, Chang LY, Bush KM, Crapo JD, Freeman BA. Detection of catalase in rat heart mitochondria. J Biol Chem 266: 22028–22034, 1991.
    Crossref | PubMed | ISI | Google Scholar
  • 542. Raineri I, Carlson EJ, Gacayan R, Carra S, Oberley TD, Huang TT, Epstein CJ. Strain-dependent high-level expression of a transgene for manganese superoxide dismutase is associated with growth retardation and decreased fertility. Free Radic Biol Med 31: 1018–1030, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 543. Rajasekaran NS, Varadharaj S, Khanderao GD, Davidson CJ, Kannan S, Firpo MA, Zweier JL, Benjamin IJ. Sustained activation of nuclear erythroid 2-related factor 2/antioxidant response element signaling promotes reductive stress in the human mutant protein aggregation cardiomyopathy in mice. Antioxid Redox Signal 14: 957–971, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 544. Ralph GS, Radcliffe PA, Day DM, Carthy JM, Leroux MA, Lee DC, Wong LF, Bilsland LG, Greensmith L, Kingsman SM, Mitrophanous KA, Mazarakis ND, Azzouz M. Silencing mutant SOD1 using RNAi protects against neurodegeneration and extends survival in an ALS model. Nat Med 11: 429–433, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 545. Ramachandran A, Lebofsky M, Weinman SA, Jaeschke H. The impact of partial manganese superoxide dismutase (SOD2)-deficiency on mitochondrial oxidant stress, DNA fragmentation and liver injury during acetaminophen hepatotoxicity. Toxicol Appl Pharmacol 251: 226–233, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 546. Ramiro-Diaz JM, Nitta CH, Maston LD, Codianni S, Giermakowska W, Resta TC, Gonzalez Bosc LV. NFAT is required for spontaneous pulmonary hypertension in superoxide dismutase 1 knockout mice. Am J Physiol Lung Cell Mol Physiol 304: L613–L625, 2013.
    Link | ISI | Google Scholar
  • 547. Ramprasath T, Murugan PS, Kalaiarasan E, Gomathi P, Rathinavel A, Selvam GS. Genetic association of Glutathione peroxidase-1 (GPx-1) and NAD(P)H:Quinone Oxidoreductase 1(NQO1) variants and their association of CAD in patients with type-2 diabetes. Mol Cell Biochem 361: 143–150, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 548. Ran Q, Liang H, Gu M, Qi W, Walter CA, Roberts LJ 2nd, Herman B, Richardson A, Van Remmen H. Transgenic mice overexpressing glutathione peroxidase 4 are protected against oxidative stress-induced apoptosis. J Biol Chem 279: 55137–55146, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 549. Ran Q, Liang H, Ikeno Y, Qi W, Prolla TA, Roberts LJ, 2nd Wolf N, Van Remmen H, Richardson A. Reduction in glutathione peroxidase 4 increases life span through increased sensitivity to apoptosis. J Gerontol A Biol Sci Med Sci 62: 932–942, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 550. Rando TA, Crowley RS, Carlson EJ, Epstein CJ, Mohapatra PK. Overexpression of copper/zinc superoxide dismutase: a novel cause of murine muscular dystrophy. Ann Neurol 44: 381–386, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 551. Ranguelova K, Ganini D, Bonini MG, London RE, Mason RP. Kinetics of the oxidation of reduced Cu,Zn-superoxide dismutase by peroxymonocarbonate. Free Radic Biol Med 53: 589–594, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 552. Raoul C, Abbas-Terki T, Bensadoun JC, Guillot S, Haase G, Szulc J, Henderson CE, Aebischer P. Lentiviral-mediated silencing of SOD1 through RNA interference retards disease onset and progression in a mouse model of ALS. Nat Med 11: 423–428, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 553. Ravn-Haren G, Olsen A, Tjonneland A, Dragsted LO, Nexo BA, Wallin H, Overvad K, Raaschou-Nielsen O, Vogel U. Associations between GPX1 Pro198Leu polymorphism, erythrocyte GPX activity, alcohol consumption and breast cancer risk in a prospective cohort study. Carcinogenesis 27: 820–825, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 554. Rayman MP. Selenium and human health. Lancet 379: 1256–1268, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 555. Rayman MP. Selenoproteins and human health: insights from epidemiological data. Biochim Biophys Acta 1790: 1533–1540, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 556. Reaume AG, Elliott JL, Hoffman EK, Kowall NW, Ferrante RJ, Siwek DF, Wilcox HM, Flood DG, Beal MF, Brown RH Jr, Scott RW, Snider WD. Motor neurons in Cu/Zn superoxide dismutase-deficient mice develop normally but exhibit enhanced cell death after axonal injury. Nat Genet 13: 43–47, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 557. Reddi AR, Culotta VC. SOD1 integrates signals from oxygen and glucose to repress respiration. Cell 152: 224–235, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 558. Ren D, Villeneuve NF, Jiang T, Wu T, Lau A, Toppin HA, Zhang DD. Brusatol enhances the efficacy of chemotherapy by inhibiting the Nrf2-mediated defense mechanism. Proc Natl Acad Sci USA 108: 1433–1438, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 559. Renko K, Werner M, Renner-Muller I, Cooper TG, Yeung CH, Hollenbach B, Scharpf M, Kohrle J, Schomburg L, Schweizer U. Hepatic selenoprotein P (SePP) expression restores selenium transport and prevents infertility and motor-incoordination in Sepp-knockout mice. Biochem J 409: 741–749, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 560. Rhee SG. Cell signaling. H2O2, a necessary evil for cell signaling. Science 312: 1882–1883, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 561. Rhee SG, Kang SW, Chang TS, Jeong W, Kim K. Peroxiredoxin, a novel family of peroxidases. IUBMB Life 52: 35–41, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 562. Rhee SG, Woo HA. Multiple functions of peroxiredoxins: peroxidases, sensors and regulators of the intracellular messenger H2O2, and protein chaperones. Antioxid Redox Signal 15: 781–794, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 563. Richters L, Lange N, Renner R, Treiber N, Ghanem A, Tiemann K, Scharffetter-Kochanek K, Bloch W, Brixius K. Exercise-induced adaptations of cardiac redox homeostasis and remodeling in heterozygous SOD2-knockout mice. J Appl Physiol 111: 1431–1440, 2011.
    Link | ISI | Google Scholar
  • 564. Ridet JL, Bensadoun JC, Deglon N, Aebischer P, Zurn AD. Lentivirus-mediated expression of glutathione peroxidase: neuroprotection in murine models of Parkinson's disease. Neurobiol Dis 21: 29–34, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 565. Rigobello MP, Callegaro MT, Barzon E, Benetti M, Bindoli A. Purification of mitochondrial thioredoxin reductase and its involvement in the redox regulation of membrane permeability. Free Radic Biol Med 24: 370–376, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 566. Rizvanov AA, Mukhamedyarov MA, Palotas A, Islamov RR. Retrogradely transported siRNA silences human mutant SOD1 in spinal cord motor neurons. Exp Brain Res 195: 1–4, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 567. Rodriguez-Iturbe B, Sepassi L, Quiroz Y, Ni Z, Wallace DC, Vaziri ND. Association of mitochondrial SOD deficiency with salt-sensitive hypertension and accelerated renal senescence. J Appl Physiol 102: 255–260, 2007.
    Link | ISI | Google Scholar
  • 568. Rong Y, Doctrow SR, Tocco G, Baudry M. EUK-134, a synthetic superoxide dismutase and catalase mimetic, prevents oxidative stress and attenuates kainate-induced neuropathology. Proc Natl Acad Sci USA 96: 9897–9902, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 569. Rosen DR, Siddique T, Patterson D, Figlewicz DA, Sapp P, Hentati A, Donaldson D, Goto J, O'Regan JP, Deng HX. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 362: 59–62, 1993.
    Crossref | PubMed | ISI | Google Scholar
  • 570. Ross AD, Banda NK, Muggli M, Arend WP. Enhancement of collagen-induced arthritis in mice genetically deficient in extracellular superoxide dismutase. Arthritis Rheum 50: 3702–3711, 2004.
    Crossref | PubMed | Google Scholar
  • 571. Rundlof AK, Arner ES. Regulation of the mammalian selenoprotein thioredoxin reductase 1 in relation to cellular phenotype, growth, and signaling events. Antioxid Redox Signal 6: 41–52, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 572. Rundlof AK, Carlsten M, Giacobini MM, Arner ES. Prominent expression of the selenoprotein thioredoxin reductase in the medullary rays of the rat kidney and thioredoxin reductase mRNA variants differing at the 5' untranslated region. Biochem J 347: 661–668, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 573. Rundlof AK, Janard M, Miranda-Vizuete A, Arner ES. Evidence for intriguingly complex transcription of human thioredoxin reductase 1. Free Radic Biol Med 36: 641–656, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 574. Sadidi M, Lentz SI, Feldman EL. Hydrogen peroxide-induced Akt phosphorylation regulates Bax activation. Biochimie 91: 577–585, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 575. Saija A, Princi P, Pisani A, Lanza M, Scalese M, Aramnejad E, Ceserani R, Costa G. Protective effect of glutathione on kainic acid-induced neuropathological changes in the rat brain. Gen Pharmacol 25: 97–102, 1994.
    Crossref | PubMed | Google Scholar
  • 576. Saito Y, Hayashi T, Tanaka A, Watanabe Y, Suzuki M, Saito E, Takahashi K. Selenoprotein P in human plasma as an extracellular phospholipid hydroperoxide glutathione peroxidase. Isolation and enzymatic characterization of human selenoprotein P. J Biol Chem 274: 2866–2871, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 577. Salmeen A, Andersen JN, Myers MP, Meng TC, Hinks JA, Tonks NK, Barford D. Redox regulation of protein tyrosine phosphatase 1B involves a sulphenyl-amide intermediate. Nature 423: 769–773, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 578. Sandbach JM, Coscun PE, Grossniklaus HE, Kokoszka JE, Newman NJ, Wallace DC. Ocular pathology in mitochondrial superoxide dismutase (Sod2)-deficient mice. Invest Ophthalmol Vis Sci 42: 2173–2178, 2001.
    PubMed | ISI | Google Scholar
  • 579. Saqib A, Prasad KM, Katwal AB, Sanders JM, Lye RJ, French BA, Annex BH. Adeno-associated virus serotype 9-mediated overexpression of extracellular superoxide dismutase improves recovery from surgical hind-limb ischemia in BALB/c mice. J Vasc Surg 54: 810–818, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 580. Schickler M, Knobler H, Avraham KB, Elroy-Stein O, Groner Y. Diminished serotonin uptake in platelets of transgenic mice with increased Cu/Zn-superoxide dismutase activity. EMBO J 8: 1385–1392, 1989.
    Crossref | PubMed | ISI | Google Scholar
  • 581. Schneider M, Forster H, Boersma A, Seiler A, Wehnes H, Sinowatz F, Neumuller C, Deutsch MJ, Walch A, Hrabe de Angelis M, Wurst W, Ursini F, Roveri A, Maleszewski M, Maiorino M, Conrad M. Mitochondrial glutathione peroxidase 4 disruption causes male infertility. FASEB J 23: 3233–3242, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 582. Schoenmakers E, Agostini M, Mitchell C, Schoenmakers N, Papp L, Rajanayagam O, Padidela R, Ceron-Gutierrez L, Doffinger R, Prevosto C, Luan J, Montano S, Lu J, Castanet M, Clemons N, Groeneveld M, Castets P, Karbaschi M, Aitken S, Dixon A, Williams J, Campi I, Blount M, Burton H, Muntoni F, O'Donovan D, Dean A, Warren A, Brierley C, Baguley D, Guicheney P, Fitzgerald R, Coles A, Gaston H, Todd P, Holmgren A, Khanna KK, Cooke M, Semple R, Halsall D, Wareham N, Schwabe J, Grasso L, Beck-Peccoz P, Ogunko A, Dattani M, Gurnell M, Chatterjee K. Mutations in the selenocysteine insertion sequence-binding protein 2 gene lead to a multisystem selenoprotein deficiency disorder in humans. J Clin Invest 120: 4220–4235, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 583. Schomburg L, Schweizer U, Holtmann B, Flohe L, Sendtner M, Kohrle J. Gene disruption discloses role of selenoprotein P in selenium delivery to target tissues. Biochem J 370: 397–402, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 584. Schriner SE, Linford NJ, Martin GM, Treuting P, Ogburn CE, Emond M, Coskun PE, Ladiges W, Wolf N, Van Remmen H, Wallace DC, Rabinovitch PS. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science 308: 1909–1911, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 585. Schriner SE, Ogburn CE, Smith AC, Newcomb TG, Ladiges WC, Dolle ME, Vijg J, Fukuchi K, Martin GM. Levels of DNA damage are unaltered in mice overexpressing human catalase in nuclei. Free Radic Biol Med 29: 664–673, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 586. Schwartz PJ, Reaume A, Scott R, Coyle JT. Effects of over- and under-expression of Cu,Zn-superoxide dismutase on the toxicity of glutamate analogs in transgenic mouse striatum. Brain Res 789: 32–39, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 587. Schweizer U, Streckfuss F, Pelt P, Carlson BA, Hatfield DL, Kohrle J, Schomburg L. Hepatically derived selenoprotein P is a key factor for kidney but not for brain selenium supply. Biochem J 386: 221–226, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 588. Seeher S, Carlson BA, Miniard AC, Wirth EK, Mahdi Y, Hatfield DL, Driscoll DM, Schweizer U. Impaired selenoprotein expression in brain triggers striatal neuronal loss leading to co-ordination defects in mice. Biochem J 462: 67–75, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 589. Seibold P, Hall P, Schoof N, Nevanlinna H, Heikkinen T, Benner A, Liu J, Schmezer P, Popanda O, Flesch-Janys D, Chang-Claude J. Polymorphisms in oxidative stress-related genes and mortality in breast cancer patients–potential differential effects by radiotherapy? Breast 22: 817–823, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 590. Seiler A, Schneider M, Forster H, Roth S, Wirth EK, Culmsee C, Plesnila N, Kremmer E, Radmark O, Wurst W, Bornkamm GW, Schweizer U, Conrad M. Glutathione peroxidase 4 senses and translates oxidative stress into 12/15-lipoxygenase dependent- and AIF-mediated cell death. Cell Metab 8: 237–248, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 591. Sengupta A, Carlson BA, Hoffmann VJ, Gladyshev VN, Hatfield DL. Loss of housekeeping selenoprotein expression in mouse liver modulates lipoprotein metabolism. Biochem Biophys Res Commun 365: 446–452, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 592. Sengupta A, Lichti UF, Carlson BA, Ryscavage AO, Gladyshev VN, Yuspa SH, Hatfield DL. Selenoproteins are essential for proper keratinocyte function and skin development. PLoS One 5: e12249, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 593. Sentman ML, Granstrom M, Jakobson H, Reaume A, Basu S, Marklund SL. Phenotypes of mice lacking extracellular superoxide dismutase and copper- and zinc-containing superoxide dismutase. J Biol Chem 281: 6904–6909, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 594. Seo MS, Kang SW, Kim K, Baines IC, Lee TH, Rhee SG. Identification of a new type of mammalian peroxiredoxin that forms an intramolecular disulfide as a reaction intermediate. J Biol Chem 275: 20346–20354, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 595. Sheldon RA, Jiang X, Francisco C, Christen S, Vexler ZS, Tauber MG, Ferriero DM. Manipulation of antioxidant pathways in neonatal murine brain. Pediatr Res 56: 656–662, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 596. Shelton P, Jaiswal AK. The transcription factor NF-E2-related factor 2 (Nrf2): a protooncogene? FASEB J 27: 414–423, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 597. Shen X, Zheng S, Metreveli NS, Epstein PN. Protection of cardiac mitochondria by overexpression of MnSOD reduces diabetic cardiomyopathy. Diabetes 55: 798–805, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 598. Sheng H, Bart RD, Oury TD, Pearlstein RD, Crapo JD, Warner DS. Mice overexpressing extracellular superoxide dismutase have increased resistance to focal cerebral ischemia. Neuroscience 88: 185–191, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 599. Sheng H, Brady TC, Pearlstein RD, Crapo JD, Warner DS. Extracellular superoxide dismutase deficiency worsens outcome from focal cerebral ischemia in the mouse. Neurosci Lett 267: 13–16, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 600. Sheng H, Kudo M, Mackensen GB, Pearlstein RD, Crapo JD, Warner DS. Mice overexpressing extracellular superoxide dismutase have increased resistance to global cerebral ischemia. Exp Neurol 163: 392–398, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 601. Sheridan PA, Zhong N, Carlson BA, Perella CM, Hatfield DL, Beck MA. Decreased selenoprotein expression alters the immune response during influenza virus infection in mice. J Nutr 137: 1466–1471, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 602. Shi M, Yang H, Motley ED, Guo Z. Overexpression of Cu/Zn-superoxide dismutase and/or catalase in mice inhibits aorta smooth muscle cell proliferation. Am J Hypertens 17: 450–456, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 603. Shi Y, Lo CS, Chenier I, Maachi H, Filep JG, Ingelfinger JR, Zhang SL, Chan JS. Overexpression of catalase prevents hypertension and tubulointerstitial fibrosis and normalization of renal angiotensin-converting enzyme-2 expression in Akita mice. Am J Physiol Renal Physiol 304: F1335–F1346, 2013.
    Link | ISI | Google Scholar
  • 604. Shi ZZ, Osei-Frimpong J, Kala G, Kala SV, Barrios RJ, Habib GM, Lukin DJ, Danney CM, Matzuk MM, Lieberman MW. Glutathione synthesis is essential for mouse development but not for cell growth in culture. Proc Natl Acad Sci USA 97: 5101–5106, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 605. Shin JH, London J, Le Pecheur M, Hoger H, Pollak D, Lubec G. Aberrant neuronal and mitochondrial proteins in hippocampus of transgenic mice overexpressing human Cu/Zn superoxide dismutase 1. Free Radic Biol Med 37: 643–653, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 606. Shin JH, London J, Le Pecheur M, Weitzdoerfer R, Hoeger H, Lubec G. Proteome analysis in hippocampus of mice overexpressing human Cu/Zn-superoxide dismutase 1. Neurochem Int 46: 641–653, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 607. Shingu M, Yoshioka K, Nobunaga M, Yoshida K. Human vascular smooth muscle cells and endothelial cells lack catalase activity and are susceptible to hydrogen peroxide. Inflammation 9: 309–320, 1985.
    Crossref | PubMed | ISI | Google Scholar
  • 608. Shrimali RK, Irons RD, Carlson BA, Sano Y, Gladyshev VN, Park JM, Hatfield DL. Selenoproteins mediate T cell immunity through an antioxidant mechanism. J Biol Chem 283: 20181–20185, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 609. Shrimali RK, Weaver JA, Miller GF, Starost MF, Carlson BA, Novoselov SV, Kumaraswamy E, Gladyshev VN, Hatfield DL. Selenoprotein expression is essential in endothelial cell development and cardiac muscle function. Neuromuscul Disord 17: 135–142, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 610. Sies H, Sharov VS, Klotz LO, Briviba K. Glutathione peroxidase protects against peroxynitrite-mediated oxidations. A new function for selenoproteins as peroxynitrite reductase. J Biol Chem 272: 27812–27817, 1997.
    Crossref | PubMed | ISI | Google Scholar
  • 611. Siklos L, Engelhardt JI, Reaume AG, Scott RW, Adalbert R, Obal I, Appel SH. Altered calcium homeostasis in spinal motoneurons but not in oculomotor neurons of SOD-1 knockout mice. Acta Neuropathol 99: 517–524, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 612. Sinet PM. Metabolism of oxygen derivatives in down's syndrome. Ann NY Acad Sci 396: 83–94, 1982.
    Crossref | PubMed | ISI | Google Scholar
  • 613. Singh A, Rangasamy T, Thimmulappa RK, Lee H, Osburn WO, Brigelius-Flohe R, Kensler TW, Yamamoto M, Biswal S. Glutathione peroxidase 2, the major cigarette smoke-inducible isoform of GPX in lungs, is regulated by Nrf2. Am J Respir Cell Mol Biol 35: 639–650, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 614. Smietana MJ, Arruda EM, Faulkner JA, Brooks SV, Larkin LM. Reactive oxygen species on bone mineral density and mechanics in Cu,Zn superoxide dismutase (Sod1) knockout mice. Biochem Biophys Res Commun 403: 149–153, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 615. Smith AD, Guidry CA, Morris VC, Levander OA. Aurothioglucose inhibits murine thioredoxin reductase activity in vivo. J Nutr 129: 194–198, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 616. Sobotta MC, Liou W, Stocker S, Talwar D, Oehler M, Ruppert T, Scharf AN, Dick TP. Peroxiredoxin-2 and STAT3 form a redox relay for H2O2 signaling. Nat Chem Biol 11: 64–70, 2015.
    Crossref | PubMed | ISI | Google Scholar
  • 617. Soerensen J, Jakupoglu C, Beck H, Forster H, Schmidt J, Schmahl W, Schweizer U, Conrad M, Brielmeier M. The role of thioredoxin reductases in brain development. PLoS One 3: e1813, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 618. Sporn MB, Liby KT. NRF2 and cancer: the good, the bad and the importance of context. Nat Rev Cancer 12: 564–571, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 619. Stralin P, Karlsson K, Johansson BO, Marklund SL. The interstitium of the human arterial wall contains very large amounts of extracellular superoxide dismutase. Arterioscler Thromb Vasc Biol 15: 2032–2036, 1995.
    Crossref | PubMed | ISI | Google Scholar
  • 620. Stranges S, Galletti F, Farinaro E, D'Elia L, Russo O, Iacone R, Capasso C, Carginale V, De Luca V, Della Valle E, Cappuccio FP, Strazzullo P. Associations of selenium status with cardiometabolic risk factors: an 8-year follow-up analysis of the Olivetti Heart study. Atherosclerosis 217: 274–278, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 621. Strassburger M, Bloch W, Sulyok S, Schuller J, Keist AF, Schmidt A, Wenk J, Peters T, Wlaschek M, Lenart J, Krieg T, Hafner M, Kumin A, Werner S, Muller W, Scharffetter-Kochanek K. Heterozygous deficiency of manganese superoxide dismutase results in severe lipid peroxidation and spontaneous apoptosis in murine myocardium in vivo. Free Radic Biol Med 38: 1458–1470, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 622. Su D, Gladyshev VN. Alternative splicing involving the thioredoxin reductase module in mammals: a glutaredoxin-containing thioredoxin reductase 1. Biochemistry 43: 12177–12188, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 623. Su D, Novoselov SV, Sun QA, Moustafa ME, Zhou Y, Oko R, Hatfield DL, Gladyshev VN. Mammalian selenoprotein thioredoxin-glutathione reductase. Roles in disulfide bond formation and sperm maturation. J Biol Chem 280: 26491–26498, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 624. Sugawara T, Lewen A, Gasche Y, Yu F, Chan PH. Overexpression of SOD1 protects vulnerable motor neurons after spinal cord injury by attenuating mitochondrial cytochrome c release. FASEB J 16: 1997–1999, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 625. Suliman HB, Ryan LK, Bishop L, Folz RJ. Prevention of influenza-induced lung injury in mice overexpressing extracellular superoxide dismutase. Am J Physiol Lung Cell Mol Physiol 280: L69–L78, 2001.
    Link | ISI | Google Scholar
  • 626. Sun QA, Kirnarsky L, Sherman S, Gladyshev VN. Selenoprotein oxidoreductase with specificity for thioredoxin and glutathione systems. Proc Natl Acad Sci USA 98: 3673–3678, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 627. Sun QA, Su D, Novoselov SV, Carlson BA, Hatfield DL, Gladyshev VN. Reaction mechanism and regulation of mammalian thioredoxin/glutathione reductase. Biochemistry 44: 14528–14537, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 628. Sun QA, Wu Y, Zappacosta F, Jeang KT, Lee BJ, Hatfield DL, Gladyshev VN. Redox regulation of cell signaling by selenocysteine in mammalian thioredoxin reductases. J Biol Chem 274: 24522–24530, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 629. Sun QA, Zappacosta F, Factor VM, Wirth PJ, Hatfield DL, Gladyshev VN. Heterogeneity within animal thioredoxin reductases. Evidence for alternative first exon splicing. J Biol Chem 276: 3106–3114, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 630. Sun R, Eriksson S, Wang L. Oxidative stress induced S-glutathionylation and proteolytic degradation of mitochondrial thymidine kinase 2. J Biol Chem 287: 24304–24312, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 631. Sunde RA, Raines AM, Barnes KM, Evenson JK. Selenium status highly regulates selenoprotein mRNA levels for only a subset of the selenoproteins in the selenoproteome. Biosci Rep 29: 329–338, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 632. Suresh A, Guedez L, Moreb J, Zucali J. Overexpression of manganese superoxide dismutase promotes survival in cell lines after doxorubicin treatment. Br J Haematol 120: 457–463, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 633. Sutton A, Nahon P, Pessayre D, Rufat P, Poire A, Ziol M, Vidaud D, Barget N, Ganne-Carrie N, Charnaux N, Trinchet JC, Gattegno L, Beaugrand M. Genetic polymorphisms in antioxidant enzymes modulate hepatic iron accumulation and hepatocellular carcinoma development in patients with alcohol-induced cirrhosis. Cancer Res 66: 2844–2852, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 634. Suvorova ES, Lucas O, Weisend CM, Rollins MF, Merrill GF, Capecchi MR, Schmidt EE. Cytoprotective Nrf2 pathway is induced in chronically txnrd 1-deficient hepatocytes. PLoS One 4: e6158, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 635. Suzuki T, Kelly VP, Motohashi H, Nakajima O, Takahashi S, Nishimura S, Yamamoto M. Deletion of the selenocysteine tRNA gene in macrophages and liver results in compensatory gene induction of cytoprotective enzymes by Nrf2. J Biol Chem 283: 2021–2030, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 636. Takagi Y, Mitsui A, Nishiyama A, Nozaki K, Sono H, Gon Y, Hashimoto N, Yodoi J. Overexpression of thioredoxin in transgenic mice attenuates focal ischemic brain damage. Proc Natl Acad Sci USA 96: 4131–4136, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 637. Takahara S. Progressive oral gangrene probably due to lack of catalase in the blood (acatalasaemia); report of nine cases. Lancet 2: 1101–1104, 1952.
    Crossref | PubMed | Google Scholar
  • 638. Takahara S, Miyamoto H. [The progressive, necrotic dental maxillitis that was considered to be the cause of the lack of catalase in the blood]. Okayama Igakkai zasshi 60: 90, 1948.
    PubMed | Google Scholar
  • 639. Takebe G, Yarimizu J, Saito Y, Hayashi T, Nakamura H, Yodoi J, Nagasawa S, Takahashi K. A comparative study on the hydroperoxide and thiol specificity of the glutathione peroxidase family and selenoprotein P. J Biol Chem 277: 41254–41258, 2002.
    Crossref | PubMed | ISI | Google Scholar
  • 640. Tan SM, Sharma A, Yuen DY, Stefanovic N, Krippner G, Mugesh G, Chai Z, de Haan JB. The modified selenenyl amide, M-hydroxy ebselen, attenuates diabetic nephropathy and diabetes-associated atherosclerosis in ApoE/GPx1 double knockout mice. PLoS One 8: e69193, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 641. Tan SM, Stefanovic N, Tan G, Wilkinson-Berka JL, de Haan JB. Lack of the antioxidant glutathione peroxidase-1 (GPx1) exacerbates retinopathy of prematurity in mice. Invest Ophthalmol Vis Sci 54: 555–562, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 642. Tanaka M, Mokhtari GK, Terry RD, Balsam LB, Lee KH, Kofidis T, Tsao PS, Robbins RC. Overexpression of human copper/zinc superoxide dismutase (SOD1) suppresses ischemia-reperfusion injury and subsequent development of graft coronary artery disease in murine cardiac grafts. Circulation 110: II200–206, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 643. Tanaka T, Nakamura H, Nishiyama A, Hosoi F, Masutani H, Wada H, Yodoi J. Redox regulation by thioredoxin superfamily: protection against oxidative stress and aging. Free Radic Res 33: 851–855, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 644. Thiels E, Urban NN, Gonzalez-Burgos GR, Kanterewicz BI, Barrionuevo G, Chu CT, Oury TD, Klann E. Impairment of long-term potentiation and associative memory in mice that overexpress extracellular superoxide dismutase. J Neurosci 20: 7631–7639, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 645. Thimmulappa RK, Mai KH, Srisuma S, Kensler TW, Yamamoto M, Biswal S. Identification of Nrf2-regulated genes induced by the chemopreventive agent sulforaphane by oligonucleotide microarray. Cancer Res 62: 5196–5203, 2002.
    PubMed | ISI | Google Scholar
  • 646. Thireau J, Poisson D, Zhang BL, Gillet L, Le Pecheur M, Andres C, London J, Babuty D. Increased heart rate variability in mice overexpressing the Cu/Zn superoxide dismutase. Free Radic Biol Med 45: 396–403, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 647. Thiruchelvam M, Prokopenko O, Cory-Slechta DA, Buckley B, Mirochnitchenko O. Overexpression of superoxide dismutase or glutathione peroxidase protects against the paraquat + maneb-induced Parkinson disease phenotype. J Biol Chem 280: 22530–22539, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 648. Tome ME, Baker AF, Powis G, Payne CM, Briehl MM. Catalase-overexpressing thymocytes are resistant to glucocorticoid-induced apoptosis and exhibit increased net tumor growth. Cancer Res 61: 2766–2773, 2001.
    PubMed | ISI | Google Scholar
  • 649. Tome ME, Lutz NW, Briehl MM. Overexpression of catalase or Bcl-2 alters glucose and energy metabolism concomitant with dexamethasone resistance. Biochim Biophys Acta 1693: 57–72, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 650. Tonks NK. Redox redux: revisiting PTPs and the control of cell signaling. Cell 121: 667–670, 2005.
    Crossref | PubMed | ISI | Google Scholar
  • 651. Torres M. Mitogen-activated protein kinase pathways in redox signaling. Front Biosci 8: d369–391, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 652. Torzewski M, Ochsenhirt V, Kleschyov AL, Oelze M, Daiber A, Li H, Rossmann H, Tsimikas S, Reifenberg K, Cheng F, Lehr HA, Blankenberg S, Forstermann U, Munzel T, Lackner KJ. Deficiency of glutathione peroxidase-1 accelerates the progression of atherosclerosis in apolipoprotein E-deficient mice. Arterioscler Thromb Vasc Biol 27: 850–857, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 653. Treiber N, Maity P, Singh K, Kohn M, Keist AF, Ferchiu F, Sante L, Frese S, Bloch W, Kreppel F, Kochanek S, Sindrilaru A, Iben S, Hogel J, Ohnmacht M, Claes LE, Ignatius A, Chung JH, Lee MJ, Kamenisch Y, Berneburg M, Nikolaus T, Braunstein K, Sperfeld AD, Ludolph AC, Briviba K, Wlaschek M, Florin L, Angel P, Scharffetter-Kochanek K. Accelerated aging phenotype in mice with conditional deficiency for mitochondrial superoxide dismutase in the connective tissue. Aging Cell 10: 239–254, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 654. Treuting PM, Linford NJ, Knoblaugh SE, Emond MJ, Morton JF, Martin GM, Rabinovitch PS, Ladiges WC. Reduction of age-associated pathology in old mice by overexpression of catalase in mitochondria. J Gerontol A Biol Sci Med Sci 63: 813–822, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 655. Trumbull KA, Beckman JS. A role for copper in the toxicity of zinc-deficient superoxide dismutase to motor neurons in amyotrophic lateral sclerosis. Antioxid Redox Signal 11: 1627–1639, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 656. Tsang CK, Liu Y, Thomas J, Zhang Y, Zheng XF. Superoxide dismutase 1 acts as a nuclear transcription factor to regulate oxidative stress resistance. Nat Commun 5: 3446, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 657. Tsuji G, Koshiba M, Nakamura H, Kosaka H, Hatachi S, Kurimoto C, Kurosaka M, Hayashi Y, Yodoi J, Kumagai S. Thioredoxin protects against joint destruction in a murine arthritis model. Free Radic Biol Med 40: 1721–1731, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 658. Tsunoda S, Kawano N, Miyado K, Kimura N, Fujii J. Impaired fertilizing ability of superoxide dismutase 1-deficient mouse sperm during in vitro fertilization. Biol Reprod 87: 121, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 659. Turanov AA, Kehr S, Marino SM, Yoo MH, Carlson BA, Hatfield DL, Gladyshev VN. Mammalian thioredoxin reductase 1: roles in redox homoeostasis and characterization of cellular targets. Biochem J 430: 285–293, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 660. Turdi S, Han X, Huff AF, Roe ND, Hu N, Gao F, Ren J. Cardiac-specific overexpression of catalase attenuates lipopolysaccharide-induced myocardial contractile dysfunction: role of autophagy. Free Radic Biol Med 53: 1327–1338, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 661. Turoczi T, Chang VW, Engelman RM, Maulik N, Ho YS, Das DK. Thioredoxin redox signaling in the ischemic heart: an insight with transgenic mice overexpressing Trx1. J Mol Cell Cardiol 35: 695–704, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 662. Uchiyama S, Shimizu T, Shirasawa T. CuZn-SOD deficiency causes ApoB degradation and induces hepatic lipid accumulation by impaired lipoprotein secretion in mice. J Biol Chem 281: 31713–31719, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 663. Ueta T, Inoue T, Furukawa T, Tamaki Y, Nakagawa Y, Imai H, Yanagi Y. Glutathione peroxidase 4 is required for maturation of photoreceptor cells. J Biol Chem 287: 7675–7682, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 664. Ufer C, Wang CC, Fahling M, Schiebel H, Thiele BJ, Billett EE, Kuhn H, Borchert A. Translational regulation of glutathione peroxidase 4 expression through guanine-rich sequence-binding factor 1 is essential for embryonic brain development. Genes Dev 22: 1838–1850, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 665. Umekawa T, Sugiyama T, Kihira T, Murabayashi N, Zhang L, Nagao K, Kamimoto Y, Ma N, Yodoi J, Sagawa N. Overexpression of thioredoxin-1 reduces oxidative stress in the placenta of transgenic mice and promotes fetal growth via glucose metabolism. Endocrinology 149: 3980–3988, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 666. Urig S, Becker K. On the potential of thioredoxin reductase inhibitors for cancer therapy. Semin Cancer Biol 16: 452–465, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 667. Ursini F, Heim S, Kiess M, Maiorino M, Roveri A, Wissing J, Flohe L. Dual function of the selenoprotein PHGPx during sperm maturation. Science 285: 1393–1396, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 668. Usui S, Oveson BC, Iwase T, Lu L, Lee SY, Jo YJ, Wu Z, Choi EY, Samulski RJ, Campochiaro PA. Overexpression of SOD in retina: need for increase in H2O2-detoxifying enzyme in same cellular compartment. Free Radic Biol Med 51: 1347–1354, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 669. Van den Bosch H, Schutgens RB, Wanders RJ, Tager JM. Biochemistry of peroxisomes. Annu Rev Biochem 61: 157–197, 1992.
    Crossref | PubMed | ISI | Google Scholar
  • 670. Van Montfort RL, Congreve M, Tisi D, Carr R, Jhoti H. Oxidation state of the active-site cysteine in protein tyrosine phosphatase 1B. Nature 423: 773–777, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 671. Van Remmen H, Williams MD, Guo Z, Estlack L, Yang H, Carlson EJ, Epstein CJ, Huang TT, Richardson A. Knockout mice heterozygous for Sod2 show alterations in cardiac mitochondrial function and apoptosis. Am J Physiol Heart Circ Physiol 281: H1422–H1432, 2001.
    Link | ISI | Google Scholar
  • 672. Van Rheen Z, Fattman C, Domarski S, Majka S, Klemm D, Stenmark KR, Nozik-Grayck E. Lung extracellular superoxide dismutase overexpression lessens bleomycin-induced pulmonary hypertension and vascular remodeling. Am J Respir Cell Mol Biol 44: 500–508, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 673. Veal EA, Day AM, Morgan BA. Hydrogen peroxide sensing and signaling. Mol Cell 26: 1–14, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 674. Veerareddy S, Cooke CL, Baker PN, Davidge ST. Gender differences in myogenic tone in superoxide dismutase knockout mouse: animal model of oxidative stress. Am J Physiol Heart Circ Physiol 287: H40–H45, 2004.
    Link | ISI | Google Scholar
  • 675. Velarde MC, Flynn JM, Day NU, Melov S, Campisi J. Mitochondrial oxidative stress caused by Sod2 deficiency promotes cellular senescence and aging phenotypes in the skin. Aging 4: 3–12, 2012.
    Crossref | PubMed | Google Scholar
  • 676. Voetsch B, Jin RC, Bierl C, Deus-Silva L, Camargo ECS, Annichino-Bizacchi JM, Handy DE, Loscalzo J. Role of promoter polymorphisms in the plasma glutathione peroxidase (GPx-3) gene as a risk factor for cerebral venous thrombosis. Stroke 39: 303–307, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 677. Vogel U, Olsen A, Wallin H, Overvad K, Tjonneland A, Nexo BA. No association between GPX Pro198Leu and risk of basal cell carcinoma. Cancer Epidemiol Biomarkers Prev 13: 1412–1413, 2004.
    PubMed | ISI | Google Scholar
  • 678. Walshe J, Serewko-Auret MM, Teakle N, Cameron S, Minto K, Smith L, Burcham PC, Russell T, Strutton G, Griffin A, Chu FF, Esworthy S, Reeve V, Saunders NA. Inactivation of glutathione peroxidase activity contributes to UV-induced squamous cell carcinoma formation. Cancer Res 67: 4751–4758, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 679. Wang F, Reece EA, Yang P. Superoxide dismutase 1 overexpression in mice abolishes maternal diabetes-induced endoplasmic reticulum stress in diabetic embryopathy. Am J Obstet Gynecol 209345.e1-7, 2013.
    ISI | Google Scholar
  • 680. Wang L, Jiang Z, Lei XG. Knockout of SOD1 alters murine hepatic glycolysis, gluconeogenesis, and lipogenesis. Free Radic Biol Med 53: 1689–1696, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 681. Wang P, Chen H, Qin H, Sankarapandi S, Becher MW, Wong PC, Zweier JL. Overexpression of human copper, zinc-superoxide dismutase (SOD1) prevents postischemic injury. Proc Natl Acad Sci USA 95: 4556–4560, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 682. Wang Q, Chen W, Bai L, Chen W, Padilla MT, Lin AS, Shi S, Wang X, Lin Y. Receptor-interacting protein 1 increases chemoresistance by maintaining inhibitor of apoptosis protein levels and reducing reactive oxygen species through a microRNA-146a-mediated catalase pathway. J Biol Chem 289: 5654–5663, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 683. Wang X, Phelan SA, Forsman-Semb K, Taylor EF, Petros C, Brown A, Lerner CP, Paigen B. Mice with targeted mutation of peroxiredoxin 6 develop normally but are susceptible to oxidative stress. J Biol Chem 278: 25179–25190, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 684. Wang X, Vatamaniuk MZ, Roneker CA, Pepper MP, Hu LG, Simmons RA, Lei XG. Knockouts of SOD1 and GPX1 exert different impacts on murine islet function and pancreatic integrity. Antioxid Redox Signal 14: 391–401, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 685. Wang X, Yun JW, Lei XG. Glutathione peroxidase mimic ebselen improves glucose-stimulated insulin secretion in murine islets. Antioxid Redox Signal 20: 191–203, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 686. Wang XD, Vatamaniuk MZ, Wang SK, Roneker CA, Simmons RA, Lei XG. Molecular mechanisms for hyperinsulinaemia induced by overproduction of selenium-dependent glutathione peroxidase-1 in mice. Diabetologia 51: 1515–1524, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 687. Wang Y, Phelan SA, Manevich Y, Feinstein SI, Fisher AB. Transgenic mice overexpressing peroxiredoxin 6 show increased resistance to lung injury in hyperoxia. Am J Respir Cell Mol Biol 34: 481–486, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 688. Watanabe R, Nakamura H, Masutani H, Yodoi J. Anti-oxidative, anti-cancer and anti-inflammatory actions by thioredoxin 1 and thioredoxin-binding protein-2. Pharmacol Ther 127: 261–270, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 689. Watson JD. Type 2 diabetes as a redox disease. Lancet 383: 841–843, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 690. Wei JP, Srinivasan C, Han H, Valentine JS, Gralla EB. Evidence for a novel role of copper-zinc superoxide dismutase in zinc metabolism. J Biol Chem 276: 44798–44803, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 691. Weisbrot-Lefkowitz M, Reuhl K, Perry B, Chan PH, Inouye M, Mirochnitchenko O. Overexpression of human glutathione peroxidase protects transgenic mice against focal cerebral ischemia/reperfusion damage. Brain Res Mol Brain Res 53: 333–338, 1998.
    Crossref | PubMed | Google Scholar
  • 692. Wen JK, Osumi T, Hashimoto T, Ogata M. Diminished synthesis of catalase due to the decrease in catalase mRNA in Japanese-type acatalasemia. Physiol Chem Phys Med NMR 20: 171–176, 1988.
    PubMed | ISI | Google Scholar
  • 693. Wen JK, Osumi T, Hashimoto T, Ogata M. Molecular analysis of human acatalasemia. Identification of a splicing mutation. J Mol Biol 211: 383–393, 1990.
    Crossref | PubMed | ISI | Google Scholar
  • 694. Wheeler MD, Nakagami M, Bradford BU, Uesugi T, Mason RP, Connor HD, Dikalova A, Kadiiska M, Thurman RG. Overexpression of manganese superoxide dismutase prevents alcohol-induced liver injury in the rat. J Biol Chem 276: 36664–36672, 2001.
    Crossref | PubMed | ISI | Google Scholar
  • 695. White CW, Avraham KB, Shanley PF, Groner Y. Transgenic mice with expression of elevated levels of copper-zinc superoxide dismutase in the lungs are resistant to pulmonary oxygen toxicity. J Clin Invest 87: 2162–2168, 1991.
    Crossref | PubMed | ISI | Google Scholar
  • 696. Widder JD, Fraccarollo D, Galuppo P, Hansen JM, Jones DP, Ertl G, Bauersachs J. Attenuation of angiotensin II-induced vascular dysfunction and hypertension by overexpression of Thioredoxin 2. Hypertension 54: 338–344, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 697. Williams MD, Van Remmen H, Conrad CC, Huang TT, Epstein CJ, Richardson A. Increased oxidative damage is correlated to altered mitochondrial function in heterozygous manganese superoxide dismutase knockout mice. J Biol Chem 273: 28510–28515, 1998.
    Crossref | PubMed | ISI | Google Scholar
  • 698. Winterbourn CC. The biological chemistry of hydrogen peroxide. Methods Enzymol 528: 3–25, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 699. Winterbourn CC, Hampton MB. Thiol chemistry and specificity in redox signaling. Free Radic Biol Med 45: 549–561, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 700. Winterbourn CC, Metodiewa D. Reactivity of biologically important thiol compounds with superoxide and hydrogen peroxide. Free Radic Biol Med 27: 322–328, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 701. Wirth EK, Bharathi BS, Hatfield D, Conrad M, Brielmeier M, Schweizer U. Cerebellar hypoplasia in mice lacking selenoprotein biosynthesis in neurons. Biol Trace Elem Res 158: 203–210, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 702. Wirth EK, Conrad M, Winterer J, Wozny C, Carlson BA, Roth S, Schmitz D, Bornkamm GW, Coppola V, Tessarollo L, Schomburg L, Kohrle J, Hatfield DL, Schweizer U. Neuronal selenoprotein expression is required for interneuron development and prevents seizures and neurodegeneration. FASEB J 24: 844–852, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 703. Wispe JR, Warner BB, Clark JC, Dey CR, Neuman J, Glasser SW, Crapo JD, Chang LY, Whitsett JA. Human Mn-superoxide dismutase in pulmonary epithelial cells of transgenic mice confers protection from oxygen injury. J Biol Chem 267: 23937–23941, 1992.
    Crossref | PubMed | ISI | Google Scholar
  • 704. Wolkart G, Kaber G, Kojda G, Brunner F. Role of endogenous hydrogen peroxide in cardiovascular ischaemia/reperfusion function: studies in mouse hearts with catalase-overexpression in the vascular endothelium. Pharmacol Res 54: 50–56, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 705. Woo HA, Yim SH, Shin DH, Kang D, Yu DY, Rhee SG. Inactivation of peroxiredoxin I by phosphorylation allows localized H2O2 accumulation for cell signaling. Cell 140: 517–528, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 706. Wood ZA, Schroder E, Robin Harris J, Poole LB. Structure, mechanism and regulation of peroxiredoxins. Trends Biochem Sci 28: 32–40, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 707. Wortmann M, Schneider M, Pircher J, Hellfritsch J, Aichler M, Vegi N, Kolle P, Kuhlencordt P, Walch A, Pohl U, Bornkamm GW, Conrad M, Beck H. Combined deficiency in glutathione peroxidase 4 and vitamin E causes multiorgan thrombus formation and early death in mice. Circ Res 113: 408–417, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 708. Woychik RP, Alagramam K. Insertional mutagenesis in transgenic mice generated by the pronuclear microinjection procedure. Int J Dev Biol 42: 1009–1017, 1998.
    PubMed | ISI | Google Scholar
  • 709. Wu CY, Steffen J, Eide DJ. Cytosolic superoxide dismutase (SOD1) is critical for tolerating the oxidative stress of zinc deficiency in yeast. PLoS One 4: e7061, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 710. Wu H, Lin L, Giblin F, Ho YS, Lou MF. Glutaredoxin 2 knockout increases sensitivity to oxidative stress in mouse lens epithelial cells. Free Radic Biol Med 51: 2108–2117, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 711. Wu J, Hecker JG, Chiamvimonvat N. Antioxidant enzyme gene transfer for ischemic diseases. Adv Drug Deliv Rev 61: 351–363, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 712. Wu S, Li Q, Du M, Li SY, Ren J. Cardiac-specific overexpression of catalase prolongs lifespan and attenuates ageing-induced cardiomyocyte contractile dysfunction and protein damage. Clin Exp Pharmacol Physiol 34: 81–87, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 713. Xia X, Zhou H, Huang Y, Xu Z. Allele-specific RNAi selectively silences mutant SOD1 and achieves significant therapeutic benefit in vivo. Neurobiol Dis 23: 578–586, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 714. Xiong Y, Liu X, Lee CP, Chua BH, Ho YS. Attenuation of doxorubicin-induced contractile and mitochondrial dysfunction in mouse heart by cellular glutathione peroxidase. Free Radic Biol Med 41: 46–55, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 715. Xiong Y, Manevich Y, Tew KD, Townsend DM. S-glutathionylation of protein disulfide isomerase regulates estrogen receptor alpha stability and function. Int J Cell Biol 2012: 273549, 2012.
    Crossref | PubMed | Google Scholar
  • 716. Xiong Y, Shie FS, Zhang J, Lee CP, Ho YS. The protective role of cellular glutathione peroxidase against trauma-induced mitochondrial dysfunction in the mouse brain. J Stroke Cerebrovasc Dis 13: 129–137, 2004.
    Crossref | PubMed | Google Scholar
  • 717. Xu B, Moritz JT, Epstein PN. Overexpression of catalase provides partial protection to transgenic mouse beta cells. Free Radic Biol Med 27: 830–837, 1999.
    Crossref | PubMed | ISI | Google Scholar
  • 718. Xu L, Emery JF, Ouyang YB, Voloboueva LA, Giffard RG. Astrocyte targeted overexpression of Hsp72 or SOD2 reduces neuronal vulnerability to forebrain ischemia. Glia 58: 1042–1049, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 719. Yamanobe T, Okada F, Iuchi Y, Onuma K, Tomita Y, Fujii J. Deterioration of ischemia/reperfusion-induced acute renal failure in SOD1-deficient mice. Free Radic Res 41: 200–207, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 720. Yan C, Huang A, Wu Z, Kaminski PM, Wolin MS, Hintze TH, Kaley G, Sun D. Increased superoxide leads to decreased flow-induced dilation in resistance arteries of Mn-SOD-deficient mice. Am J Physiol Heart Circ Physiol 288: H2225–H2231, 2005.
    Link | ISI | Google Scholar
  • 721. Yan X, Pepper MP, Vatamaniuk MZ, Roneker CA, Li L, Lei XG. Dietary selenium deficiency partially rescues type 2 diabetes-like phenotypes of glutathione peroxidase-1-overexpressing male mice. J Nutr 142: 1975–1982, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 722. Yang H, Roberts LJ, Shi MJ, Zhou LC, Ballard BR, Richardson A, Guo ZM. Retardation of atherosclerosis by overexpression of catalase or both Cu/Zn-superoxide dismutase and catalase in mice lacking apolipoprotein E. Circ Res 95: 1075–1081, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 723. Yang H, Shi M, VanRemmen H, Chen X, Vijg J, Richardson A, Guo Z. Reduction of pressor response to vasoconstrictor agents by overexpression of catalase in mice. Am J Hypertens 16: 1–5, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 724. Yang H, Zhou L, Wang Z, Roberts LJ 2nd, Lin X, Zhao Y, Guo Z. Overexpression of antioxidant enzymes in ApoE-deficient mice suppresses benzo(a)pyrene-accelerated atherosclerosis. Atherosclerosis 207: 51–58, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 725. Yant LJ, Ran Q, Rao L, Van Remmen H, Shibatani T, Belter JG, Motta L, Richardson A, Prolla TA. The selenoprotein GPX4 is essential for mouse development and protects from radiation and oxidative damage insults. Free Radic Biol Med 34: 496–502, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 726. Yatmaz S, Seow HJ, Gualano RC, Wong ZX, Stambas J, Selemidis S, Crack PJ, Bozinovski S, Anderson GP, Vlahos R. Glutathione peroxidase-1 reduces influenza A virus-induced lung inflammation. Am J Respir Cell Mol Biol 48: 17–26, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 727. Yen HC, Oberley TD, Vichitbandha S, Ho YS, St Clair DK. The protective role of manganese superoxide dismutase against adriamycin-induced acute cardiac toxicity in transgenic mice. J Clin Invest 98: 1253–1260, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 728. Yim MB, Chock PB, Stadtman ER. Copper, zinc superoxide dismutase catalyzes hydroxyl radical production from hydrogen peroxide. Proc Natl Acad Sci USA 87: 5006–5010, 1990.
    Crossref | PubMed | ISI | Google Scholar
  • 729. Ying W, Anderson CM, Chen Y, Stein BA, Fahlman CS, Copin JC, Chan PH, Swanson RA. Differing effects of copper,zinc superoxide dismutase overexpression on neurotoxicity elicited by nitric oxide, reactive oxygen species, and excitotoxins. J Cereb Blood Flow Metab 20: 359–368, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 730. Yoo MH, Xu XM, Carlson BA, Gladyshev VN, Hatfield DL. Thioredoxin reductase 1 deficiency reverses tumor phenotype and tumorigenicity of lung carcinoma cells. J Biol Chem 281: 13005–13008, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 731. Yoshida T, Maulik N, Engelman RM, Ho YS, Das DK. Targeted disruption of the mouse Sod1 gene makes the hearts vulnerable to ischemic reperfusion injury. Circ Res 86: 264–269, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 732. Yoshida T, Maulik N, Engelman RM, Ho YS, Magnenat JL, Rousou JA, Flack JE 3rd, Deaton D, Das DK. Glutathione peroxidase knockout mice are susceptible to myocardial ischemia reperfusion injury. Circulation 96: II-216-220, 1997.
    ISI | Google Scholar
  • 733. Yoshida T, Watanabe M, Engelman DT, Engelman RM, Schley JA, Maulik N, Ho YS, Oberley TD, Das DK. Transgenic mice overexpressing glutathione peroxidase are resistant to myocardial ischemia reperfusion injury. J Mol Cell Cardiol 28: 1759–1767, 1996.
    Crossref | PubMed | ISI | Google Scholar
  • 734. Yoshihara D, Fujiwara N, Kato S, Sakiyama H, Eguchi H, Suzuki K. Alterations in renal iron metabolism caused by a copper/zinc-superoxide dismutase deficiency. Free Radic Res 46: 750–757, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 735. Yoshihara E, Masaki S, Matsuo Y, Chen Z, Tian H, Yodoi J. Thioredoxin/Txnip: redoxisome, as a redox switch for the pathogenesis of diseases. Front Immunol 4: 514, 2014.
    Crossref | PubMed | ISI | Google Scholar
  • 736. Yu DH, Yi JK, Yuh HS, Park S, Kim HJ, Bae KB, Ji YR, Kim NR, Park SJ, Kim do H, Kim SH, Kim MO, Lee JW, Ryoo ZY. Over-expression of extracellular superoxide dismutase in mouse synovial tissue attenuates the inflammatory arthritis. Exp Mol Med 44: 529–535, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 737. Yu F, Sugawara T, Nishi T, Liu J, Chan PH. Overexpression of SOD1 in transgenic rats attenuates nuclear translocation of endonuclease G and apoptosis after spinal cord injury. J Neurotrauma 23: 595–603, 2006.
    Crossref | PubMed | ISI | Google Scholar
  • 738. Yue TL, McKenna PJ, Ruffolo RR Jr, Feuerstein G. Carvedilol, a new beta-adrenoceptor antagonist and vasodilator antihypertensive drug, inhibits superoxide release from human neutrophils. Eur J Pharmacol 214: 277–280, 1992.
    Crossref | PubMed | ISI | Google Scholar
  • 739. Zaghloul N, Nasim M, Patel H, Codipilly C, Marambaud P, Dewey S, Schiffer WK, Ahmed M. Overexpression of extracellular superoxide dismutase has a protective role against hyperoxia-induced brain injury in neonatal mice. FEBS J 279: 871–881, 2012.
    Crossref | PubMed | ISI | Google Scholar
  • 740. Zanetti M, Katusic ZS, O'Brien T. Adenoviral-mediated overexpression of catalase inhibits endothelial cell proliferation. Am J Physiol Heart Circ Physiol 283: H2620–H2626, 2002.
    Link | ISI | Google Scholar
  • 741. Zhang B, Wang Y, Su Y. Peroxiredoxins, a novel target in cancer radiotherapy. Cancer Lett 286: 154–160, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 742. Zhang DD. The Nrf2-Keap1-ARE signaling pathway: The regulation and dual function of Nrf2 in cancer. Antioxid Redox Signal 13: 1623–1626, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 743. Zhang DD, Lo SC, Cross JV, Templeton DJ, Hannink M. Keap1 is a redox-regulated substrate adaptor protein for a Cul3-dependent ubiquitin ligase complex. Mol Cell Biol 24: 10941–10953, 2004.
    Crossref | PubMed | ISI | Google Scholar
  • 744. Zhang H, Go YM, Jones DP. Mitochondrial thioredoxin-2/peroxiredoxin-3 system functions in parallel with mitochondrial GSH system in protection against oxidative stress. Arch Biochem Biophys 465: 119–126, 2007.
    Crossref | PubMed | ISI | Google Scholar
  • 745. Zhang H, Joseph J, Felix C, Kalyanaraman B. Bicarbonate enhances the hydroxylation, nitration, and peroxidation reactions catalyzed by copper, zinc superoxide dismutase. Intermediacy of carbonate anion radical. J Biol Chem 275: 14038–14045, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 746. 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. J Neuropathol Exp Neurol 59: 53–61, 2000.
    Crossref | PubMed | ISI | Google Scholar
  • 747. Zhang M, Dong Y, Xu J, Xie Z, Wu Y, Song P, Guzman M, Wu J, Zou MH. Thromboxane receptor activates the AMP-activated protein kinase in vascular smooth muscle cells via hydrogen peroxide. Circ Res 102: 328–337, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 748. Zhang X, Dong F, Li Q, Borgerding AJ, Klein AL, Ren J. Cardiac overexpression of catalase antagonizes ADH-associated contractile depression and stress signaling after acute ethanol exposure in murine myocytes. J Appl Physiol 99: 2246–2254, 2005.
    Link | ISI | Google Scholar
  • 749. Zhang X, Klein AL, Alberle NS, Norby FL, Ren BH, Duan J, Ren J. Cardiac-specific overexpression of catalase rescues ventricular myocytes from ethanol-induced cardiac contractile defect. J Mol Cell Cardiol 35: 645–652, 2003.
    Crossref | PubMed | ISI | Google Scholar
  • 750. Zhang Y, Ikeno Y, Qi W, Chaudhuri A, Li Y, Bokov A, Thorpe SR, Baynes JW, Epstein C, Richardson A, Van Remmen H. Mice deficient in both Mn superoxide dismutase and glutathione peroxidase-1 have increased oxidative damage and a greater incidence of pathology but no reduction in longevity. J Gerontol A Biol Sci Med Sci 64: 1212–1220, 2009.
    Crossref | PubMed | ISI | Google Scholar
  • 751. Zhao CR, Gao ZH, Qu XJ. Nrf2-ARE signaling pathway and natural products for cancer chemoprevention. Cancer Epidemiol 34: 523–533, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 752. Zhao H, Kim G, Liu C, Levine RL. Transgenic mice overexpressing methionine sulfoxide reductase A: characterization of embryonic fibroblasts. Free Radic Biol Med 49: 641–648, 2010.
    Crossref | PubMed | ISI | Google Scholar
  • 753. Zhou J, Huang K, Lei XG. Selenium and diabetes–evidence from animal studies. Free Radic Biol Med 65: 1548–1556, 2013.
    Crossref | PubMed | ISI | Google Scholar
  • 754. Zhu JH, Lei XG. Double null of selenium-glutathione peroxidase-1 and copper, zinc-superoxide dismutase enhances resistance of mouse primary hepatocytes to acetaminophen toxicity. Exp Biol Med 231: 545–552, 2006.
    Crossref | ISI | Google Scholar
  • 755. Zhu JH, Lei XG. Lipopolysaccharide-induced hepatic oxidative injury is not potentiated by knockout of GPX1 and SOD1 in mice. Biochem Biophys Res Commun 404: 559–563, 2011.
    Crossref | PubMed | ISI | Google Scholar
  • 756. Zhu JH, McClung JP, Zhang X, Aregullin M, Chen C, Gonzalez FJ, Kim TW, Lei XG. Comparative impacts of knockouts of two antioxidant enzymes on acetaminophen-induced hepatotoxicity in mice. Exp Biol Med 234: 1477–1483, 2009.
    Crossref | ISI | Google Scholar
  • 757. Zhu JH, Zhang X, McClung JP, Lei XG. Impact of Cu,Zn-superoxide dismutase and Se-dependent glutathione peroxidase-1 knockouts on acetaminophen-induced cell death and related signaling in murine liver. Exp Biol Med 231: 1726–1732, 2006.
    Crossref | ISI | Google Scholar
  • 758. Zhu JH, Zhang X, Roneker CA, McClung JP, Zhang S, Thannhauser TW, Ripoll DR, Sun Q, Lei XG. Role of copper,zinc-superoxide dismutase in catalyzing nitrotyrosine formation in murine liver. Free Radic Biol Med 45: 611–618, 2008.
    Crossref | PubMed | ISI | Google Scholar
  • 759. Zmijewski JW, Lorne E, Zhao X, Tsuruta Y, Sha Y, Liu G, Abraham E. Antiinflammatory effects of hydrogen peroxide in neutrophil activation and acute lung injury. Am J Respir Crit Care Med 179: 694–704, 2009.
    Crossref | PubMed | ISI | Google Scholar