Widget HTML Atas

A Royal Elf Of Abalon Anna Dye Pdf Free Download

I. INTRODUCTION

A.  From Oxidative Damage to Redox Regulation: Historic Background

The presence of free radicals in biological materials was discovered less than 50 years ago (114). Soon therafter, Denham Harman hypothesized that oxygen radicals may be formed as by-products of enzymic reactions in vivo. In 1956, he described free radicals as a Pandora's box of evils that may account for gross cellular damage, mutagenesis, cancer, and, last but not least, the degenerative process of biological aging (234,235).

The science of free radicals in living organisms entered a second era after McCord and Fridovich (386) discovered the enzyme superoxide dismutase (SOD) and, finally, convinced most colleagues that free radicals are important in biology. Numerous researchers were now inspired to investigate oxidative damage inflicted by radicals upon DNA, proteins, lipids, and other components of the cell (reviewed in Ref. 49).

A third era began with the first reports describing advantageous biological effects of free radicals. Mittal and Murard (394) provided suggestive evidence that the superoxide anion (O 2 ·), through its derivative, the hydroxyl radical, stimulates the activation of guanylate cyclase and formation of the "second messenger" cGMP. Similar effects were reported for the superoxide derivative hydrogen peroxide (615). Ignarro and Kadowitz (272) and Moncada and colleagues (456) discovered independently the role of nitric oxide (NO) as a regulatory molecule in the control of smooth muscle relaxation and in the inhibition of platelet adhesion. Roth and Dröge (472) found that in activated T cells the superoxide anion or low micromolar concentrations of hydrogen peroxide increase the production of the T-cell growth factor interleukin-2, an immunologically important T-cell protein. Keyse and Tyrrell (300) showed that hydrogen peroxide induces the expression of the heme oxygenase (HO-1) gene. Storz and colleagues (551) reported the induction of various genes in bacteria by hydrogen peroxide, and Schreck and Baeuerle (501) reported the activation of the transcription factor nuclear factor κB (NF-κB) by hydrogen peroxide in mammalian cells.

At the beginning of the 21st century, there is now a large body of evidence showing that living organisms have not only adapted to an unfriendly coexistence with free radicals but have, in fact, developed mechanisms for the advantageous use of free radicals. Important physiological functions that involve free radicals or their derivatives include the following: regulation of vascular tone, sensing of oxygen tension and regulation of functions that are controlled by oxygen concentration, enhancement of signal transduction from various membrane receptors including the antigen receptor of lymphocytes, and oxidative stress responses that ensure the maintenance of redox homeostasis (see Table 1). The field of redox regulation is also receiving growing attention from clinical colleagues in view of the role that oxidative stress has been found to play in numerous disease conditions. These pathological conditions demonstrate the biological relevance of redox regulation. The delicate balance between the advantageous and detrimental effects of free radicals is clearly an important aspect of life. The science of biological "redox regulation" is a rapidly growing field of research that has impact on diverse disciplines including physiology, cell biology, and clinical medicine.

Table 1. Important physiological functions that involve free radicals or their derivatives

Type of Radical Source of Radical Physiological Process
Nitric oxide (NO·) Nitric oxide synthase Smooth muscle relaxation (control of vascular tone) and various other cGMP-dependent functions
Superoxide (O 2  · ) and related ROS NAD(P)H oxidase Control of ventilationControl of erythropoietin production and other hypoxia-inducible functionsSmooth muscle relaxationSignal transduction from various membrane receptors/enhancement of immunological functions
Superoxide (O 2  · ) and related ROS Any source Oxidative stress responses and the maintenance of redox homeostasis

B.  About This Review

We are now living in a particularly exciting time of redox research where information from different fields and independent approaches is falling into place and beginning to reveal a meaningful picture. This is a good time for a broad overview that summarizes the main principles of redox regulation. In textbook style, this review describes the current knowledge and paradigms but does not discuss future research directions, historical controversies, or experimental models. Moreover, it was not within the scope of this review to deal with all the details. Even the more than 600 references cited here do not cover all relevant publications in the field. For the interested reader, a number of more detailed and specific reviews on this topic are recommended (see Refs. 13, 20, 45, 66, 122, 125, 183, 211, 212,251, 294, 333, 337, 397, 446, 506, 510, 512, 632, 659).

II. MAJOR TYPES OF FREE RADICALS AND THEIR DERIVATIVES IN LIVING ORGANISMS

A.  Reactive Oxygen Species

The superoxide anion is formed by the univalent reduction of triplet-state molecular oxygen (3O2). This process is mediated by enzymes such as NAD(P)H oxidases and xanthine oxidase or nonenzymically by redox-reactive compounds such as the semi-ubiquinone compound of the mitochondrial electron transport chain (see Fig. 1). SODs convert superoxide enzymically into hydrogen peroxide (130,187). In biological tissues superoxide can also be converted nonenzymically into the nonradical species hydrogen peroxide and singlet oxygen (1O2) (549). In the presence of reduced transition metals (e.g., ferrous or cuprous ions), hydrogen peroxide can be converted into the highly reactive hydroxyl radical (·OH) (101). Alternatively, hydrogen peroxide may be converted into water by the enzymes catalase or glutathione peroxidase (Fig. 1). In the glutathione peroxidase reaction glutathione is oxidized to glutathione disulfide, which can be converted back to glutathione by glutathione reductase in an NADPH-consuming process (Fig. 1).

Fig. 1.

Fig. 1.Pathways of reactive oxygen species (ROS) production and clearance. GSH, glutathione; GSSG, glutathione disulfide.

Because superoxide and NO are readily converted by enzymes or nonenzymic chemical reactions into reactive nonradical species such as singlet oxygen (1O2), hydrogen peroxide, or peroxynitrite (ONOO), i.e., species which can in turn give rise to new radicals, the regulatory effects of these nonradical species have also been included in this review. Most of the regulatory effects are indeed not directly mediated by superoxide but rather by its reactive oxygen species (ROS) derivatives. Frequently, different reactive species coexist in the reactive environment and make it difficult to identify unequivocally which agent is responsible for a given biological effect.

B.  Reactive Nitrogen Species

The NO radical (NO·) is produced in higher organisms by the oxidation of one of the terminal guanido-nitrogen atoms ofl-arginine (437). This process is catalyzed by the enzyme NOS. Depending on the microenvironment, NO can be converted to various other reactive nitrogen species (RNS) such as nitrosonium cation (NO+), nitroxyl anion (NO) or peroxynitrite (ONOO) (546). Some of the physiological effects may be mediated through the intermediate formation of S-nitroso-cysteine orS-nitroso-glutathione (207).

C.  Key Message From Section ii

The most relevant radicals in biological regulation are superoxide and NO (see Table 1). These radicals are formed by two groups of enzymes, i.e., the NAD(P)H oxidase and NOS isoforms, respectively. Many regulatory effects are mediated by hydrogen peroxide and other ROS that are chemically derived from superoxide.

III. OXIDATIVE STRESS RESPONSE AS A MODEL OF REDOX SIGNALING

A.  Maintenance of "Redox Homeostasis"

The term redox signaling is widely used to describe a regulatory process in which the signal is delivered through redox chemistry. Redox signaling is used by a wide range of organisms, including bacteria, to induce protective responses against oxidative damage and to reset the original state of "redox homeostasis" after temporary exposure to ROS.

1.  Oxidant-antioxidant balance

Free radicals and reactive nonradical species derived from radicals exist in biological cells and tissues at low but measurable concentrations (228, 527). Their concentrations are determined by the balance between their rates of production and their rates of clearance by various antioxidant compounds and enzymes, as illustrated schematically in Figure2. Halliwell and Gutteridge (228) have defined antioxidants as substances that are able, at relatively low concentrations, to compete with other oxidizable substrates and, thus, to significantly delay or inhibit the oxidation of these substrates. This definition includes the enzymes SOD, glutathione peroxidase (GPx), and catalase, as well as nonenzymic compounds such as α-tocopherol (vitamin E), β-carotene, ascorbate (vitamin C), and glutathione.

Fig. 2.

Fig. 2.Mechanisms of redox homeostasis. Balance between ROS production and various types of scavengers. The steady-state levels of ROS are determined by the rate of ROS production and their clearance by scavenging mechanisms. Certain antioxidative enzymes including superoxide dismutase (SOD), glutathione peroxidase, catalase, and thioredoxin are potent ROS scavengers but occur in cells only at relatively low concentrations. The same is true for nonenzymic antioxidants. Amino acids and proteins are also ROS scavengers. Amino acids are less effective than the classical antioxidants on a molar basis, but their cumulative intracellular concentration is >0.1 M.

In addition, there are compounds that have a relatively low specific antioxidative activity, i.e., on a molar basis, but, when present at high concentrations, can contribute significantly to the overall ROS scavenging activity. The most prominent examples of such high-level, low-efficiency antioxidants are free amino acids, peptides, and proteins. Practically all amino acids can serve as targets for oxidative attack by ROS, although some amino acids such as tryptophan, tyrosine, histidine, and cysteine are particularly sensitive to ROS (126, 128,545). Because the cumulative intracellular concentration of free amino acids is on the order of 10−1 M, free amino acids are quantitatively important ROS scavengers (see sect.iii B5).

2.  Oxidized proteins as substrates for proteolytic digestion and their contribution to redox homeostasis

Oxygen radicals and other ROS cause modifications of proteins (reviewed in Ref. 220). These oxidative modifications may lead to changes in protein function, chemical fragmentation, or increased susceptibility to proteolytic attack (124,543, 631). Proteolytic degradation is executed mainly by proteasomes (219). In one of the studies, proteolysis was estimated to increase more than 11-fold after exposure to superoxide or hydrogen peroxide (127). Proteolysis is enhanced by 20–400 μM hydrogen peroxide, whereas millimolar concentrations inhibit proteolysis and may lead to the intracellular accumulation of oxidized proteins (220,544).

The proteins may differ strongly in their susceptibility to oxidative damage. The redox-sensitive amino acids of bovine serum albumin, for example, were shown to be oxidized about twice as fast as those of glutamine synthase (54), and intact proteins are less sensitive to oxidation than misfolded proteins (161). These findings implicate that 1) phylogenetic evolution has selected for protein structures that are relatively well-protected against oxidation and 2) ROS scavenging activities of intact proteins are weaker than those of misfolded proteins or equivalent concentrations of their constituent amino acids. Protein oxidation and enhanced proteolytic degradation cause, therefore, a net increase in ROS scavenging capacity as schematically illustrated in Figure 2. Preliminary experiments showed that treatment of human skeletal muscle cells with proteasome inhibitors causes a substantial increase in intracellular ROS levels and that this increase is reversed by the addition of free amino acids (R. Breitkreutz and W. Dröge, unpublished observations). More systematic studies are needed to determine the relative contribution of proteins, free amino acids, and classical antioxidant compounds and enzymes to the total ROS scavenging capacity of different cells and tissues.

3.  Changes in oxidant-antioxidant balance as a trigger for redox regulation: the theory of redox homeostasis and the existence of different quasi-stable states

Living cells and tissues have several mechanisms for reestablishing the original redox state after a temporary exposure to increased ROS or RNS concentrations. The production of NO (NO·), for example, is subject to direct feedback inhibition of NOS by NO (see sect. iii B1). Elevated ROS concentrations induce in many cells the expression of genes whose products exhibit antioxidative activity (Fig. 2). A major mechanism of redox homeostasis is based on the ROS-mediated induction of redox-sensitive signal cascades that lead to increased expression of antioxidative enzymes or an increase in the cystine transport system, which, in turn, facilitates in certain cell types the increase in intracellular glutathione (see Fig. 2; for details see sect. iii,B2–B4). Moreover, because proteins generally provide less ROS scavenging activity than an equivalent amount of the free amino acids contained in them, it is reasonable to assume that oxidative enhancement of proteolysis also contributes, at least to some extent, to the maintenance of redox homeostasis (see Fig. 2 and sect.iii B5).

Cells or tissues are in a stable state if the rates of ROS production and scavenging capacity are essentially constant and in balance (see Fig. 2 and baseline level in Fig. 3). Redox signaling requires that this balance is disturbed, either by an increase in ROS concentrations or a decrease in the activity of one or more antioxidant systems (Fig. 3). In higher organisms, such an oxidative event may be induced in a regulated fashion by the activation of endogenous RNS- or ROS-generating systems (see sect.iv A). However, similar responses may be induced by oxidative stress conditions generated by environmental factors (see sect. iii A4). If the initial increase in ROS is relatively small, the antioxidative response may be sufficient to compensate for the increase in ROS and to reset the original balance between ROS production and ROS scavenging capacity. Thus physiological manifestations of redox regulation involve typically a temporary increase and/or a temporary shift of the intracellular thiol/disulfide redox state toward more oxidative conditions, as illustrated in Figure3. In the long run, these mechanisms tend to maintain a stable state called redox homeostasis.

Fig. 3.

Fig. 3.Regulatory events and their dysregulation depend on the magnitude and duration of the change in ROS or reactive nitrogen species (RNS) concentration. ROS and RNS normally occur in living tissues at relatively low steady-state levels. The regulated increase in superoxide or nitric oxide production leads to a temporary imbalance that forms the basis of redox regulation. The persistent production of abnormally large amounts of ROS or RNS, however, may lead to persistent changes in signal transduction and gene expression, which, in turn, may give rise to pathological conditions.

Under certain conditions, however, ROS production is increased more strongly and persistently, and the antioxidative response may not be sufficient to reset the system to the original level of redox homeostasis. In such cases, the system may still reach an equilibrium according to the model in Figure 2, but the resulting quasi-stable state may now be associated with higher ROS concentrations and different levels of free amino acids and/or different patterns of gene expression due to redox-sensitive signaling pathways. Indications for such a shift to more oxidative conditions have been seen in the process of aging (see sect. vi C), implying that a pro-oxidative shift may not always be associated with an overtly pathological condition.

Pathological conditions (see sect. vi, H–P) may develop in more extreme cases of persistently high ROS levels (see Fig.3). Again, these conditions do not necessarily involve a loss of homeostasis but rather a chronic shift in the level of homeostasis. Accordingly, pathological symptoms may result from both the damaging effects of ROS and from ROS-mediated changes in gene expression.

4.  Redox regulation by changes in the thiol/disulfide redox state

In several cases (see Table 2), changes in the intracellular thiol/disulfide redox state have been shown to trigger the same redox-responsive signaling proteins and pathways as those triggered by hydrogen peroxide (28,192, 248, 323). Bacterial OxyR (see sect. iii B2) is a model case of a redox-sensitive signaling protein that may be activated either directly by hydrogen peroxide or, alternatively, by changes in the intracellular glutathione redox state (28). Conversely, protein tyrosine phosphatases (see sect. v B) are inactivated either by ROS or by pro-oxidative changes in the intracellular thiol/disulfide redox state. Accordingly, this review also addresses the role of the thiol/disulfide redox state in the control of cell function.

Table 2. Signaling mechanisms that respond to changes in the thiol/disulfide redox state

Example Reference No. Section
AP-1 transcription factor in human T cells 192 v H
NF-κB transcription factor in human T cells 192 v I
yAP-1 transcription factor in S. cerevisiae 323 iii B3
Control of K+ channel activity in the carotid body 9 iv C
Human insulin receptor kinase activity 494 v C
Bacterial OxyR 28 iii B2
Protein tyrosine phosphatases 43 v B
Src family kinases 248 v D
JNK and p38 MAPK signaling pathways 248 v E
Amplification of immunologic functions 248 iv F
Signaling in replicative senescence 536 vi D

5.  Effects of carcinogens and allergens on redox-responsive signaling pathways

Certain alkylating agents play an important role as environmental carcinogens by exerting effects on redox-sensitive signaling pathways similar to those induced by ROS (621). There is a strong possibility that the resulting dysregulation of signaling cascades may contribute to the process of carcinogenesis (see sect.vi I).

Certain heavy metal compounds are important environmental allergens. A point in case is the sulfhydryl-reactive compound mercury dichloride (HgCl2). Treatment of murine lymphocytes with HgCl2 was found to induce strong aggregation and activation of the protein tyrosine kinase p56lck. These findings suggest that the oxidation of redox-reactive sulfhydryl groups of certain signaling proteins leads to a dysregulation of lymphocytes and may contribute thereby to the development of allergies (411).

These examples illustrate that the mechanisms of redox regulation may be targets for hazardous environmental agents. However, a more detailed discussion of this point would exceed the scope of this review.

B.  Examples of Redox Signaling in the Maintenance of Redox Homeostasis

The maintenance of redox homeostasis involves regulatory mechanisms that are capable of sensing NO or ROS. NO, for example, inhibits directly the NO-producing enzyme. ROS stimulate in many cells the expression of compensatory gene products. Some but not all of these regulatory mechanisms are well characterized at the molecular level.

1.  NO-mediated feedback inhibition of NOS

Many of the enzymes that utilize a heme prosthetic group in catalysis are inactivated by NO. This applies to the heme-containing cytochrome P-450-related enzyme NOS, leading to feedback inhibition of NO production by NO (6,79, 215). The inhibition of NOS was shown to involve the formation of a ferrous-nitrosyl complex (6).

2.  Oxidative stress-inducible gene expression in bacteria as a model for ROS-responsive signaling pathways

The protective responses of bacteria against ROS are among the best investigated examples of redox-regulated gene expression (reviewed in Ref. 658). Prokaryotes have several different signaling pathways for responding to ROS or to alterations in the intracellular redox state (28, 45, 176,551, 634, 657). Studies inEscherichia coli revealed that low levels of ROS activate the expression of several gene products involved in the antioxidant defense including Mn-SOD (238), catalase (649), hydroperoxidase I (katG), an alkylhydoperoxide reductase (ahpCF), glutathione reductase (gorA), glutaredoxin 1 (grxA), and a regulatory RNA (oxyS).

At least nine proteins that are synthesized in Salmonella typhimurium and E. coli after exposure to hydrogen peroxide are under the control of the oxyR locus (45,108, 138). The OxyR protein controls protective responses against normally lethal doses of hydrogen peroxide or against killing by heat. In addition, it negatively autoregulates its own expression (109, 489). Hydrogen peroxide does not stimulate the synthesis of OxyR but converts the reduced form of OxyR into its oxidized and regulatory competent form (Fig. 4) (45,551, 657). Exposure of the OxyR protein to hydrogen peroxide results initially in the conversion of Cys-199 to a sulfenic acid derivative which subsequently forms an intramolecular disulfide bond with Cys-208, as illustrated schematically in Figure 4(657). Because the redox potential of OxyR is −185 mV, and the thiol-disulfide redox state of the normal bacterial cytoplasm is typically −260 to −280 mV, the OxyR is normally in the reduced form and exemplifies proteins that can be activated either through direct oxidation by hydrogen peroxide or by an oxidative shift in the thiol/disulfide redox status, as illustrated in Figure 4(28). Both oxidized and reduced OxyR are able to bind to the oxyR-oxyS promoter region but exhibit different binding characteristics (578). The formation of disulfide bonds can be reversed by glutaredoxin 1 and by thioredoxin (Trx) (657). Because OxyR controls the induction of glutathione reductase and glutaredoxin 1, the OxyR response is part of an autoregulatory circuit.

Fig. 4.

Fig. 4.Schematic model of OxyR activation. The regulatory protein OxyR is activated by the formation of disulfide bridges. This process is mediated either by hydrogen peroxide (H2O2) or by oxidative changes in the intracellular thiol/disulfide redox state.

Protective responses against superoxide are controlled by the sox locus (255, 581), which regulates the induction of ∼10 proteins, including Mn-SOD, NADPH:ferredoxin oxidoreductase, and glucose-6-phosphate dehydrogenase (reviewed in Ref. 45). One of the gene products from the sox locus, the SoxR protein, exists in solution as a homodimer containing two stable (2Fe-2S) centers that are anchored to four cysteine residues near the COOH terminals (67,255, 634). Under normal physiological conditions, these iron-sulfur centers are in the reduced state. They are readily oxidized, however, under oxidative stress (149). The oxidative process is reversible if the oxidative stress conditions are removed. Only the oxidized form of SoxR stimulates transcription of soxS (67, 149,195). Both the oxidized and the reduced forms of SoxR can bind to DNA but interact differently with RNA polymerase.

The redox-sensitive proteins OxyR and SoxR have been described here as two particularly prominent and well-investigated cases of bacterial redox regulation. Additional examples of redox-responsive regulatory mechanisms in prokaryotic cells have been reviewed by Bauer et al. (45).

3.  The oxidative stress response of the budding yeast Saccharomyces cerevisiae

Treatment of S. cerevisiae with hydrogen peroxide activates the yAP-1 transcription factor that binds specifically to the AP-1 site of the eukaryotic AP-1 family of transcription factors (323). Similar activation can be induced by diamide or diethyl maleate, i.e., two thiol oxidants that modulate the intracellular reduced glutathione (GSH) state. The yAP-1 transcription factor is involved in protective responses against oxidative stress, inducing Trx production from the TRX2 gene (323). Overexpression of yAP-1 on a multicopy plasmid was also found to increase the expression of SOD, glutathione reductase, and glucose-6-phosphate dehydrogenase, whereas yAP-1 mutant strains showed greatly decreased levels of these enzymes (498). Oxidative activation of yAP-1 operates at the posttranslational level and involves the translocalization of yAP-1 from the cytoplasm to the nucleus. This translocation is controlled by a cysteine-rich domain at the COOH terminus. Three conserved cysteine residues in this region are believed to be important in sensing the redox state (324). The transcription factor SKN7 cooperates with yAP-1 and is, therefore, also critically involved in oxidative stress response (399).

The response of S. cerevisiae to hydrogen peroxide (0.4 mM for 15 min) was studied by two-dimensional gel electrophoresis of total cell proteins, revealing an increase in the expression of 115 proteins and a decrease in the expression of another 52 proteins (200). The induced proteins include Mn-SOD, Cu/Zn-SOD, glutathione reductase, catalase, Trx reductase, cytochrome-cperoxidase, several proteasome subunits, and heat shock proteins. In addition, carbohydrate metabolism is rapidly redirected to the regeneration of NADPH at the expense of glycolysis.

4.  Protective responses in higher organisms

The redox-responsive signal cascades in mammalian cells show certain similarities to the oxidative stress response of S. cerevisiae. While activated macrophages and neutrophils in inflamed tissue of higher organisms generate massive amounts of ROS to kill environmental pathogens (see sect. iv A2), other host cells must be protected against this oxidative burst. Lymphocytes recruited into the inflammatory environment initiate antigen-specific immunological effector mechanisms and can function only because they are able to activate powerful protective mechanisms against the oxidative stress. Exposure to ROS or changes in the intracellular thiol redox state modulate various signal transduction cascades and increase the activities of several transcription factors (reviewed in Refs. 12, 13, 183, 337, 397, 446, 506, 512).

The oxidoreductase Trx is one of the proteins that is inducibly expressed in lymphocytes and other cells by hydrogen peroxide, ultraviolet (UV) irradiation, and other conditions of oxidative stress (373, 480, 565). Together with the glutathione system, Trx plays a key role in the maintenance of a reducing intracellular redox state in higher organisms. The 5′-upstream sequence of the human Trx gene contains putative binding sites for the redox-responsive transcription factors AP-1 and NF-κB (407, 408).

Exposure of macrophages to low levels of ROS or other inducers of oxidative stress induces the expression of peroxyredoxin I (i.e., a Trx peroxidase), heme oxygenase-1 (HO-1), and the cystine transporter x c (278, 486). Because the plasma concentration of reduced cysteine is relatively low, the cystine transporter plays a limiting role in the cellular supply of cyst(e)ine and in the biosynthesis of glutathione in macrophages and lymphocytes (reviewed in Ref. 160). ROS and various other oxidants were also found to induce MnSOD mRNA levels to a moderate extent in several cell types (522).

The redox control of the heme oxygenase-1 (HO-1) gene is one of the best studied models of redox regulation in mammalian cells. HO-1 induction in skin fibroblasts may serve as an inducible defense pathway to remove heme liberated by oxidants. The HO-1 protein and mRNA are strongly induced by ROS, physiological doses of UVA irradiation, and various other inducers of oxidative stress, including NO (237, 300, 585,586). The UVA response is synergistically enhanced by depletion of intracellular glutathione. The sustained induction of HO-1 mRNA and its inducibility in many tissues and various mammalian species has rendered HO-1 mRNA a useful marker for cellular oxidative stress at the mRNA level. In murine macrophages, HO-1 expression is induced by hydrogen peroxide via AP-1 (16, 91,92). Activation of ERK and p38 MAPK was implicated in HO-1 expression in chicken hepatoma cells (167).

C.  Key Message From Section iii

Free radicals and their derivatives exist in living tissues at low but measurable concentrations that are determined by the balance between the rates of radical production and their corresponding rates of clearance.

The relatively high intracellular concentrations of glutathione and other antioxidative compounds provide a strong basal scavenging capacity.

The ROS-mediated oxidation of proteins and the resulting increase in proteolytic degradation is expected to cause an increase in intracellular ROS scavenging capacity and may thereby contribute to the maintenance of redox homeostasis.

A wide variety of living organisms including bacteria have the capacity to respond to increased levels of ROS with an increase in intracellular glutathione or with increased expression of proteins/enzymes with ROS scavenging capacity. This process is known as "oxidative stress response." The resulting increase in ROS clearance capacity allows cells and tissues to maintain redox homeostasis.

The inducibility of HO-1 mRNA in many tissues and various mammalian species has rendered HO-1 mRNA a useful marker for cellular oxidative stress at the mRNA level.

The regulation of physiological responses by free radicals (see sects.iv and v) is embedded in these basic mechanisms of redox homeostasis.

Oxidative stress responses provide some of the best studied examples of redox-responsive signaling pathways.

Changes in the intracellular thiol/disulfide redox state cause in many cases chemical modifications of redox-sensitive signaling compounds similar to the modifications caused by ROS. The corresponding signal pathways may become sensitive to systemic changes in the thiol/disulfide redox state (see sect. vi, C andH).

IV. NITRIC OXIDE AND REACTIVE OXYGEN SPECIES AS REGULATORY MEDIATORS OF PHYSIOLOGICAL RESPONSES

The following sections describe cases of redox regulation that serve physiological functions other than protection against oxidative stress and redox homeostasis. These cases typically involve the regulated production of NO or ROS and the resulting effects on defined signaling cascades.

A.  Regulated Production of Free Radicals in Higher Organisms

1.  Regulated production of NO

The enzyme NOS exists in three isoforms, i.e., neuronal NOS (nNOS; type I) (69), inducible NOS (iNOS; type II) (639), and endothelial NOS (eNOS; type III) (330). Many tissues express one or more of these isoforms. The isoforms nNOS and eNOS are constitutively expressed, but their activity is regulated by the intracellular calcium concentration. The isoform iNOS is inducibly expressed in macrophages after stimulation by cytokines, lipopolysaccharides, and other immunologically relevant agents (reviewed in Ref. 59). Expression of iNOS is regulated at the transcriptional and posttranscriptional level by signaling pathways that involve agents such as the redox-responsive transcription factor NF-κB or mitogen-activated protein kinases (MAPKs) (366). The rate of NO synthesis is also determined to some extent by the availability of the substrate l-arginine and by the cofactor tetrahydrobiopterin (BH4).

2.  ROS production by phagocytic NADPH oxidase: the oxidative burst

Activated macrophages and neutrophils can produce large amounts of superoxide and its derivatives via the phagocytic isoform of NADPH oxidase. This enzyme is a heme-containing protein complex illustrated schematically in Figure 5. In an inflammatory environment hydrogen peroxide is produced by activated macrophages at an estimated rate of 2–6 × 10−14mol·h−1·cell−1 and may reach a concentration of 10–100 μM in the vicinity of these cells (299, 335, 412). The massive production of antimicrobial and tumoricidal ROS in an inflammatory environment is called the "oxidative burst" and plays an important role as a first line of defense against environmental pathogens. The physiological relevance of NADPH oxidase as a defense agent is suggested by the observation that mice lacking the NADPH oxidase components gp91phox or p47 exhibit reduced resistance to infection (132, 148, 169,400, 448, 466,521). The combined activities of NADPH oxidase and myeloperoxidase in phagocytes leads, in addition, to the production of hypochlorous acid (HClO), one of the strongest physiological oxidants and a powerful antimicrobial agent (229,500). Stimulated neutrophils and macrophages generate also singlet oxygen by reactions that involve either myeloperoxidase or NADPH oxidase (549). Importantly, however, physiologically relevant ROS concentrations can also modulate redox-sensitive signal cascades and enhance immunological functions of lymphocytes (see sect. iv F).

Fig. 5.

Fig. 5.Structure of neutrophil NAD(P)H oxidase. The enzyme consists of the membrane-bound cytochrome b 558 complex comprising gp91phox and p22phox, the cytosolic proteins p47 and p67, and a low-molecular-weight G protein of the rac family.

Phagocytic NADPH oxidase becomes activated upon translocation of cytosolic p47, p67, and a G protein of the rac family to the membrane-bound cytochrome b 558 complex that contains gp91phox and p22 (Fig. 5). The catalytic moiety gp91phox is a plasma membrane-associated complex protein containing a flavin-adenine dinucleotide component and two hemes (for details see Refs. 211, 333). The activation of phagocytic NADPH oxidase can be induced by microbial products such as bacterial lipopolysaccharide, by lipoproteins, or by cytokines such as interferon-γ, interleukin-1β, or interleukin-8 (60). The activation of NADPH oxidase is mainly controlled by the rac isoform rac2 in neutrophils and rac1 in macrophages and monocytes (567, 346).

3.  ROS production by NAD(P)H oxidases in nonphagocytic cells

The production of ROS by nonphagocytic NAD(P)H oxidase isoforms plays a role in the regulation of intracellular signaling cascades in various types of nonphagocytic cells including fibroblasts, endothelial cells, vascular smooth muscle cells, cardiac myocytes, and thyroid tissue (36, 163, 210,211, 288-290, 354,387, 553, 557, 559,571, 572, 660). In most of these cases, rac1 is involved in the induction of NAD(P)H oxidase activity (289, 290, 660). Muscle cells and fibroblasts account for most of the superoxide produced in the normal vessel wall.

The NAD(P)H oxidase isoforms of the cardiovascular system are membrane-associated enzymes that appear to utilize both NADH and NADPH (211). The rate of superoxide production in nonphagocytic cells is only about one-third of that of neutrophils. Vascular smooth muscle cells, in contrast to neutrophils, endothelial cells, or fibroblasts, generate superoxide and hydrogen peroxide mainly intracellularly.

The cardiovascular NAD(P)H oxidase isoforms are induced by hormones, hemodynamic forces, or by local metabolic changes (211). Angiotensin II increases NAD(P)H-driven superoxide production in cultured vascular smooth muscle cells and fibroblasts. Thrombin, platelet-derived growth factor (PDGF), and tumor necrosis factor-α (TNF-α) stimulate NAD(P)H oxidase-dependent superoxide production in vascular smooth muscle cells. Interleukin-1, TNF-α, and platelet-activating factor increase NAD(P)H-dependent superoxide production in fibroblasts. Mechanical forces stimulate NAD(P)H oxidase activity in endothelial cells. Reoxygenation stimulates NAD(P)H oxidase activity in cardiac myocytes.

Whereas the gp91phox of phagocytic cells (see Fig. 5) has also been found in endothelial cells, other nonphagocytic cells appear to utilize structural homologs of gp91phox. Several novel homologs of gp91phox have recently been identified and are the subject of intense research. The gp91phox homolog p138tox is (the catalytic unit of) the NADPH oxidase that supports thyroid hormone biosynthesis (163). An NAD(P)H oxidase with low affinity for oxygen and high affinity for cyanide is believed to act as one of the sensors for oxygen tension in the carotid body; these sensors control the rate of ventilation (8). The function of oxygen sensing is apparently shared by several proteins, including a nonmitochondrial cytochromeb 558, a mitochondrial protein, and possibly a third heme protein (329, 659). A similar group of proteins was suggested to be involved as oxygen sensors in the regulation of erythropoietin production in human hepatoma cells (659). A microsomal NADH oxidase was implicated as an oxygen sensor in bovine pulmonary and coronary arteries, where changes in oxygen tension regulate vascular relaxation through changes in superoxide production and cGMP formation (632).

There is a strong possibility that rac-like proteins also occur in plants (5, 624), where they may be involved in the induction of NAD(P)H oxidase-like enzymes (569). The oxidative burst in plants is an effective bactericidal mechanism.

4.  ROS production in lymphocytes by 5-lipoxygenase

The enzyme 5-lipoxygenase (5-LO) has been identified as an inducible source of ROS production in lymphocytes (60,359), but the evidence for its physiological role in redox signaling is still scarce. Lipoxygenases are non-heme-containing dioxygenases that oxidize polyunsaturated fatty acids at specific carbon sites to give hydroperoxy fatty acid derivatives with conjugated double bonds. The numbers in specific enzyme names such as 5-LO, 12-LO, or 15-LO refer to the arachidonic acid site that is predominantly oxidized (644). 5-LO is best known for its role in the biosynthesis of the leukotrienes A4, B4, C4, D4, and E4. The oxidized metabolites generated by 5-LO were found to change the intracellular redox balance and to induce signal transduction pathways and gene expression. 5-LO was shown to be involved in the production of hydrogen peroxide by T lymphocytes after ligation of the CD28 costimulatory receptor (359) and in response to interleukin-1β (60). Hydrogen peroxide production in response to CD28 stimulation was found to be decreased by specific inhibitors of 5-LO or lipid peroxidation but not by inhibitors of NADPH oxidase or cyclooxygenase. The physiological role of 5-LO in CD28-mediated signal transduction remains to be confirmed by more detailed studies.

5.  ROS production by cyclooxygenase

Cyclooxygenase-1 has been implicated in ROS production in cells stimulated with TNF-α, interleukin-1, bacterial lipopolysaccharide, or the tumor promoter 4-O-tetradecanoylphorbol-13-acetate (TPA) (178, 406). A role for cyclooxygenase in the inflammatory response of 5-LO-defective mice was also suggested (206); however, the evidence for the participation of cyclooxygenase in redox signaling is still scarce.

B.  Regulation of Vascular Tone and Other Regulatory Functions of NO

The most prominent cases of NO-mediated regulation are in the control of vascular tone and platelet adhesion. The mechanism of these processes is particularly well characterized at the molecular level.

1.  Signals involving guanylate cyclase

Guanylate cyclase belongs to the family of heterodimeric heme proteins and catalyzes the formation of cGMP, which is utilized as an intracellular amplifier and second messenger in a large range of physiological responses (197). NO binds to the heme moiety of guanylate cyclase, disrupting the planar form of the heme iron. The resulting conformational change activates the enzyme (273). Its product cGMP modulates the function of protein kinases, phosphodiesterases, ion channels, and other physiologically important targets (497). The most important examples include the regulation of smooth muscle tone (272) and the inhibition of platelet adhesion (456). Vascular smooth muscle relaxation is mediated by a cGMP-dependent protein kinase that phosphorylates and activates a calcium-sensitive potassium channel (27). Other examples have been reviewed by Deora and Lander (140).

Superoxide and hydrogen peroxide may also play a role in the activation of guanylate cyclase (82, 394,615). Since micromolar concentrations of hydrogen peroxide stimulate NOS (190), there is a strong possibility that the activation of guanylate cyclase by superoxide and hydrogen peroxide may be mediated, at least in part, indirectly via NOS. Other studies suggest that compound I, a form of catalase, may play a critical role in the activation of guanylate cyclase by hydrogen peroxide (82). After interacting with hydrogen peroxide, catalase is converted into an oxidized heme intermediate called compound I, which is normally converted back into its reduced form by a second molecule of hydrogen peroxide.

2.  RNS-mediated switch of function: aconitase and iron-regulatory protein-1

In iron-sulfur proteins, iron is bound simultaneously to inorganic sulfide groups and cysteine thiolate groups of the proteins. Such proteins are sensitive to both ROS and RNS. Oxidation usually results in dissolution of the iron-sulfur cluster and loss of function (85, 97, 250).

A special case is mammalian (4Fe-4S) aconitase, which is involved in the citric acid cycle. RNS inhibit the aconitase activity by disrupting the Fe-S clusters but expose simultaneously an RNA-binding site with specificity for the iron-response elements of the transferrin receptor and ferritin mRNAs. In this form the protein is called iron-regulatory protein-1 and is involved in iron homeostasis. In this case, the RNS-induced loss of one enzyme function is associated with the gain of or switch to another function for the same protein.

3.  NO-mediated activation of the GTP-binding protein p21 Ras and protein kinase cascades

In human peripheral blood mononuclear cells and endothelial cells, NO was found to activate all three MAPK pathways (89,336, 339, 445). The effect has been attributed to the NO-mediated stimulation of a membrane-associated protein tyrosine phosphatase activity which may lead to the dephosphorylation and activation of the Src family protein tyrosine kinase p56lck (339). Another Src family protein kinase, p60c-src , was also found to be activated by NO in fibroblasts or immunoprecipitates (15). The activation was associated with autophosphorylation at Tyr-416 and S-S bond-mediated aggregation of the kinase molecules. NO may also activate Ras byS-nitrosylation of cysteine-118 (335-338).

C.  ROS Formation as a Sensor for Changes in Oxygen Concentration: Control of Ventilation

Oxygen homeostasis is maintained in higher organisms by a tight regulation of the red blood cell mass and respiratory ventilation. Carotid bodies are sensory organs that detect changes in arterial blood oxygen. They are composed of glomus type I chemoreceptor cells that release neurotransmitters in response to hypoxia. This process changes the level of electrical activity in the efferent fibers of the carotid sinus nerve, thus relaying the sensory information to the brain stem neurons that regulate breathing. A growing body of evidence indicates that changes in oxygen concentration are sensed independently by several different ROS-producing proteins including a b-type cytochrome with properties similar to those of the cytochromeb 558 in the NADPH oxidase complex in neutrophils (reviewed in Ref. 7). Other studies suggest that changes in the rate of mitochondrial ROS production may play a major role in oxygen sensing by the carotid body (80, 449). Furthermore, it was found that the β-subunit of the potassium channel resembles the structure of NADPH-oxidoreductase (221). It is widely accepted that transduction of the sinus nerve signal involves changes in the K+-channel conductivity of type I cells in response to changes in oxygen tension (356). Exactly how changes in ROS production are translated into changes in K+ channel activity remains to be established. Acker and Xue (9) implicated ROS-mediated changes in the glutathione redox state in the control of K+ efflux and the corresponding Ca2+ influx.

D.  The Oxygen Sensor in the Regulation of Erythropoietin Production: Redox Regulation Through the Transcription Factor Hypoxia-Inducible Factor 1

The red blood cell mass is regulated by the hormone erythropoietin, which is mainly produced by kidney and liver cells, following stimulation by hypoxia. The oxygen-sensing mechanisms are still unclear and the subject of controversy. It is clear, however, that changes in oxygen tension are sensed by changes in ROS production (174, 264, 291,413). Expression of erythropoietin protein or mRNA in hepatoma cell lines or perfused kidneys was found to be strongly repressed by hydrogen peroxide. Treatment of normoxic cells with exogenous catalase stimulated erythropoietin production (94). A cytochrome b-like NAD(P)H oxidase is believed to play a role as a major oxygen sensor and ROS producer. However, the role of a mitochondrial mechanism of ROS production has also been reported (7, 103, 173,659).

The erythropoietin gene is controlled by the transcription factor hypoxia-inducible factor 1 (HIF-1) (608) (as illustrated schematically in Fig. 6). HIF-1 is a heterodimeric protein composed of the subunits HIF-1α and HIF-1β (Fig. 6). The HIF-1α and HIF-1β genes are constitutively expressed, and changes in oxygen tension fail to affect the concentration of the HIF-1β subunit. In contrast, under normoxic conditions, HIF-1α is rapidly degraded by proteasomes in an ROS-dependent manner (265). Hypoxia decreases the ROS-mediated degradation of HIF-1α and enhances thereby the formation of the heterodimeric complex (Fig. 6) (510,659).

Fig. 6.

Fig. 6.Regulation of the transcription factor hypoxia-inducible factor 1 (HIF-1). HIF-1 is a heterodimeric protein composed of the subunits HIF-1α and HIF-1β, the genes of which are both constitutively expressed. Changes in oxygen tension fail to affect the concentration of HIF-1β, whereas HIF-1α is rapidly degraded under normoxic conditions by proteasomes in a ROS-dependent fashion. In these cells, ROS production is tightly linked to oxygen concentrations and, therefore, serves as a sensor for oxygen tension. The figure illustrates ROS production by a membrane-bound NAD(P)H oxidase.

HIF-1α and HIF-1β mRNAs are expressed in most if not all human and rodent tissues (510). The number of target genes activated by HIF-1 continues to increase, and targets include genes whose protein products are involved in angiogenesis, energy metabolism, erythropoiesis, cell proliferation and viability, vascular remodeling, and vasomotor responses (reviewed in Ref. 510).

The same transcription factor is now known to control the production of a variety of hypoxia-regulated hormones and proteins including the vascular endothelial growth factor (VEGF) that stimulates the formation of new blood vessels (80) and the tyrosine hydroxylase (TH) that facilitates the control of ventilation by the carotid body (reviewed in Refs. 510, 659).

E.  Redox Regulation of Cell Adhesion

Controlled changes in the adhesive properties of cells and tissues play an important role in many biological processes. Adhesion of leukocytes to endothelial cells in postcapillary venules, for example, is an early step in chronic inflammation and depends on the expression of cell-surface receptors known as cell adhesion molecules (17). Cell adhesion molecules are also implicated in embryogenesis, cell growth, differentiation, and wound repair (186). The expression of cell adhesion molecules is stimulated by bacterial lipopolysaccharides and by various cytokines such as TNF, interleukin-1α, and interleukin-1β (17).

The adherence of leukocytes to endothelial cells is also induced by ROS (478, 509). This effect is abolished by catalase but not by superoxide dismutase, suggesting that hydrogen peroxide and not superoxide is the effective agent (509). Moreover, the oxidant-induced adherence of neutrophils is inhibited by hydroxyl radical scavengers or iron chelators, suggesting that the induction of adherence may be mediated by hydroxyl radicals generated from hydrogen peroxide within the cell.

Adhesion of neutrophils to endothelial cells involves the intercellular adhesion molecule-1 (ICAM-1), CD11b/CD18, and L-selectin (509). In addition, ROS treatment of endothelial cells induces the phosphorylation of the focal adhesion kinase pp125FAK, a cytosolic tyrosine kinase that has been implicated in the oxidant-mediated adhesion process (477, 488).

F.  Redox-Mediated Amplification of Immune Responses

Lymphocytes are the carriers of immunological specificity and, therefore, play an important role in the defense against environmental pathogens. A sophisticated combination of regulatory mechanisms ensures that even minute amounts of pathogen activate highly aggressive responses without causing major damage to host tissue. The immune response typically involves the lymphocyte receptor for antigen, receptors for costimulatory signals, and various types of cytokine (507, 516). The response is also subject to regulation by redox processes. The functional activation of T lymphocytes is strongly enhanced by ROS and/or by a shift in the intracellular glutathione redox state (248). Superoxide and/or physiologically relevant concentrations of hydrogen peroxide were shown to augment the production of interleukin-2 by antigenically or mitogenically stimulated T cells in various experimental systems (358, 472). Low micromolar concentrations of hydrogen peroxide were also shown to induce the expression of the interleukin-2 receptor in a mouse T-cell lymphoma line (358).

T-cell functions such as interleukin-2 production can be readily induced in vitro without the addition of ROS simply by exposure of T cells to relatively high concentrations of antigen or antibodies that bind to the T-cell receptor and the costimulatory receptor CD28 (352). In T cells, strong activation of the costimulatory receptor CD28 causes a significant decrease in intracellular glutathione levels and the endogenous production of hydrogen peroxide (359). An intact organism, however, is typically infected by relatively small concentrations of pathogens, and ligand concentrations for the T-cell receptor and the costimulatory CD28 receptor are expected to be suboptimal, at least in the initial phases of the infection.

Exposure of T lymphocytes to physiologically relevant concentrations of environmental ROS or other inducers of moderate oxidative stress does not bypass the requirement for signaling cascades initiated by specific cell membrane receptors, but such exposure can amplify signaling cascades after relatively weak receptor stimulation (248). Signaling cascades from the various cell membrane receptors are differentially regulated by ROS. Transcription from the interleukin-2 promoter is strongly enhanced in Jurkat T cells by exposure to 50 μM hydrogen peroxide in combination with but not without anti-CD28 ligands, indicating that the redox effect can enhance the stimulatory signal from the antigen receptor but cannot replace the signal from the CD28 costimulatory receptor. Thus hydrogen peroxide from the inflammatory environment thus appears to decrease the triggering thresholds of the antigen receptor-dependent signal cascades as illustrated schematically in Figure7. It is often critical for the survival of the infected host that a specific immune response be induced before optimal antigen doses have accumulated.

Fig. 7.

Fig. 7.Functions of ROS in the immunological response against environmental pathogens. The massive production of ROS (oxidative burst) by activated macrophages in the inflammatory environment provides a first line of defense against environmental pathogens. A certain fraction of pathogens, however, may escape this rapid but moderately effective manifestation of "innate immunity" and may generate within a few days a large progeny of pathogens. Antigenic peptides generated within the activated macrophages by the breakdown of pathogens are presented by major histocompatibility complex (MHC) determinants to the antigen receptors (AR) of T lymphocytes. This interaction triggers the proliferation and differentiation of the T cells and leads within a few days to a large progeny of immunological effector cells. The effector cells provide a highly effective and antigen-specific immunological defense. ROS that are concomitantly produced by the activated macrophages in the inflammatory environment enhance the AR-mediated signal cascades and decrease thereby the activation threshold of the T cells. Without this effect, the T lymphocytes would require relatively large concentrations of antigenic peptides and would lose valuable time in their "race" with the proliferating pathogens. In this situation time may be a matter of life or death for the organism.

The importance of the oxidative microenvironment in the activation of immune responses by small concentrations of antigen in vivo is exemplified by studies of the generation of cytotoxic T lymphocytes in mice after in vivo immunization with small numbers of syngeneic tumor cells or with cells expressing foreign minor histocompatibility antigens (474). The simultaneous injection of glutathione was found to enhance immunization with relatively large numbers of stimulator cells but suppress in vivo immunization by small numbers of stimulator cells (i.e., small amounts of antigen). The glutathione-dependent enhancement of the response to high doses of antigen is in line with the requirement for reducing conditions in the execution phase of the signaling cascade and for the DNA synthesis (see sect. v K) and suggests that this requirement may become a limiting factor under conditions of high antigen concentrations.

Last but not least, there is evidence that the intracellular redox state also modulates the immunological functions of macrophages. Hamuro et al. (230) reported that macrophages vary strongly in their release of prostaglandins, interleukin-6, and interleukin-12, depending on the intracellular content of glutathione. The balance between "reductive" and "oxidative" macrophages regulates thereby the ratio of helper T cells of type 1 versus type 2 (TH1/TH2).

G.  Role of ROS in Programmed Cell Death

1.  Induction and execution of apoptosis

Apoptosis is a special form of programmed cell death that plays an indispensable role in the development and homeostasis of multicellular organisms (636). An increase in cellular ROS production is often observed in apoptotic processes triggered by various stimuli including APO-1/Fas/CD95 ligands (41,171, 257, 287, 317,588, 623, 655). However, some authors found that triggering of the APO-1/Fas/CD95 receptor does not induce ROS production (268), and some authors observed membrane changes typical of apoptosis in the absence of ROS (96, 281), indicating that pro-oxidative conditions are not a general prerequisite for apoptotic cell death. Nevertheless, high ROS concentrations induce apoptotic cell death in various cell types (reviewed in Refs. 162, 535), suggesting that ROS contribute to cell death whenever they are generated in the context of the apoptotic process. Exposure of T lymphocytes to relatively moderate concentrations of hydrogen peroxide was found to induce a CD95-independent apoptotic process that requires mitochondrial ROS production and the activation of NF-κB (162). This underscores the need to transfer lymphocytes to more reducing conditions for the development of immunological effector functions (see sect. v K).

2.  NO-dependent apoptosis

NO-dependent apoptosis has been observed in several experimental models and certain clinical pathologies (10,18, 76, 132, 575). Induction of apoptosis by NO is associated with a decrease in the concentration of cardiolipin, decreased activity of the mitochondrial electron transport chain, and release of mitochondrial cytochrome c into the cytosol (589,592). However, some cell types, such as endothelial cells from the microvasculature, are extremely resistant to the induction of apoptosis by NO (357), and low concentrations of NO provide protection from apoptotic cell death in various cell types by inhibiting certain caspases (113, 304,349, 479). High intracellular glutathione levels are associated with increased resistance to NO-mediated apoptosis (589).

3.  Induction of cell death by TNF-α

TNF-α induces cell death in many types of tumor cells and has been used in model systems for studies of the molecular mechanisms of cell death. In transformed cell lines TNF-α induces endogenous ROS production by mitochondria (423, 504). Whether and how these ROS contribute to the induction of cell death depends on the signaling and execution pathways that are activated (145). In leukocytes and fibroblasts, TNF-α induces the release of superoxide by the activation of membrane-bound NADPH oxidases. This process induces proliferation or cell death depending on the condition of the ROS-producing cell (249,310, 387, 504,515).

H.  Regulatory Role of ROS in Plants

The attempted infection of plants by an avirulent pathogen was found to induce a protective "hypersensitive response" (HR) that involves the generation of superoxide and hydrogen peroxide (331). These ROS induce a number of defense genes or drive cells into apoptosis (279, 331,347). A study of the infection of Arabidopsisleaves by avirulent Pseudomonas syringae revealed that the local primary oxidative burst leads to subsequent secondary microbursts in distant leaves and contributes thereby to the development of a state of systemic acquired resistance (SAR) (21). Activation of SAR is associated with the expression of several gene families, i.e., the pathogenesis-related (PR) genes (587). Salicylic acid acts as a signaling molecule in the system (617). The receptor for salicylic acid has catalase activity (105). Binding of salicylic acid causes the downregulation of catalase activity and leads thereby to a significant increase in the intracellular hydrogen peroxide concentration at sites distant to the original insult (105). The distant accumulation of (secondary) low levels of hydrogen peroxide induces the expression of various defense genes and gene products including chitinase and peroxidase (633). Two stress MAPKs ofArabidopsis, i.e., AtMPK3 and AtMPK6, are activated by hydrogen peroxide through the MAPK kinase ANP1 (318).

I.  Key Messages From Section iv

Numerous physiological functions are controlled by redox-responsive signaling pathways. These cases of redox regulation typically involve the regulated production of NO or ROS by NOS or NAD(P)H oxidase, respectively, and the effects of these compounds on specific signaling cascades.

The redox-sensitive target molecules of these signaling cascades and their chemical modification by oxidative agents are, in most cases, not yet well understood. A few well-studied examples show that for a given redox-responsive signaling protein NO or ROS may induce a gain of function, a loss of function, or a switch from one function to another.

The surprisingly large number of NAD(P)H oxidase isoforms and NO synthetases is by itself a strong indication for the physiological relevance of redox signaling in biological regulation.

A particularly active field of research deals with the role of ROS in the regulation of cardiovascular functions. Practically all cell types in the vascular wall produce ROS. Upon stimulation by growth factors or cytokines, vascular NAD(P)H oxidases produce superoxide and other ROS that activate multiple intracellular signaling pathways. Thus ROS play a decisive role in the normal functioning of cardiac and vascular cells (211, 559).

The regulation of vascular tone by cGMP is a special case. The enzyme guanylate cyclase was found to be activated by both NO and hydrogen peroxide (272, 394, 615,632). This effect of hydrogen peroxide is believed to play a role in the oxygen-dependent regulation of cGMP levels and vascular contraction.

Other responses to changes in oxygen tension include the control of ventilation by the carotid body and the regulated production of certain hormones such as erythropoietin, VEGF, and insulin-like growth factor II (IGF-II). All of these hormones are under transcriptional control of the transcription factor HIF-1 (510,659).

One of the well-studied cell types with respect to redox-responsive signaling cascades is the lymphoid cell. The redox sensitivity of antigen-induced signaling cascades strongly suggests that massive ROS production by phagocytic cells in the inflammatory environment (i.e., the oxidative burst) serves not only as a first line of defense against environmental pathogens but also enhances the response of lymphocytes to (small amounts of) antigen.

V. REDOX-SENSITIVE TARGETS IN SIGNALING CASCADES

A.  Role of ROS in Receptor-Mediated Signaling Pathways: the EGF Receptor as a Case in Point

There are various examples of growth factors, cytokines, or other ligands that trigger ROS production in nonphagocytic cells through their corresponding membrane receptors (see sect.iv A3). Such ROS production can mediate a positive feedback effect on signal transduction from these receptors since intracellular signaling is often enhanced by ROS or by a pro-oxidative shift of the intracellular thiol/disulfide redox state, as illustrated in Figure 8(36, 39, 50, 115,301, 354, 355, 525,554, 557, 558, 571,606, 635). The molecular details of the oxidative enhancement are not entirely clear. However, work on different receptors and signaling pathways revealed certain consistent patterns that have important physiological implications.

Fig. 8.

Fig. 8.Role of ROS in epidermal growth factor (EGF) receptor-mediated signaling. The interaction of the EGF receptor or other related membrane receptors with corresponding ligands leads to activation of NAD(P)H oxidase and the production of superoxide and hydrogen peroxide. Hydrogen peroxide, in turn, facilitates the autophosphorylation of the membrane receptor and the induction of the signal cascade. The activation of NAD(P)H oxidase by other membrane receptors such as the angiotensin II receptor can provide a cooperative effect that contributes to the autophosphorylation of the EGF receptor and to the activation of the EGF-dependent signaling cascade.

For example, the role of ROS has been demonstrated for nerve growth factor (NGF) signaling in neuronal cells (558), for epidermal growth factor (EGF) signaling in human epidermoid carcinoma cells (36), and for PDGF (37,246). Stimulation by any of these growth factors results in a transient increase in intracellular ROS through the signaling protein Rac1. Elimination of hydrogen peroxide by catalase was shown to inhibit EGF- and NGF-induced tyrosine phosphorylation of various cellular proteins, including phosphorylation of the growth factor receptor itself (see Fig. 8). In the case of the EGF receptor, induction of ROS production was reported to require the kinase activity of the receptor but not the phosphorylation of the four autophosphorylation sites at the COOH terminus of the EGF receptor (36). This unexpected finding remains to be confirmed. In view of the many growth factors, cytokines, or other ligands that trigger endogenous production of ROS, there is a strong possibility that the redox dependency of the signal transduction process may facilitate synergistic interactions between different types of membrane receptors, as illustrated in Figure 8. This physiologically advantageous cooperativity between different receptors may have been a major driving force in the phylogenetic evolution of the rather complex mechanism of redox regulation. The enhancing effect of ROS on tyrosine phosphorylation and catalytic activation, as exemplified by the EGF receptor, applies in a similar manner to various other protein components of intracellular signaling pathways.

A particularly well-studied case of cooperativity is the interaction between the angiotensin II type 1 receptor and the EGF receptor or PDGF-β receptor. The angiotensin II type 1 receptor is a G protein-coupled receptor that mediates growth effects in vascular smooth muscle cells, cardiomyocytes, and cardiac fibroblasts. In addition, this receptor was shown to mediate responses that are normally activated by tyrosine kinase-linked receptors as exemplified by those for EGF and PDGF. The transactivation of these two growth factor receptors by angiotensin II is mediated by ROS and inhibited by antioxidants such as N-acetylcysteine (212, 246, 458,607). Preliminary evidence suggests that angiotensin II-mediated transactivation of the EGF receptor involves the intermittant activation of the Src family kinase p60c-src by hydrogen peroxide (591), which can directly activate p60c-src in mouse fibroblasts (4; see sect.v D).

B.  Enhancement of Signaling Cascades by Oxidative Inhibition of Protein Tyrosine Phosphatases

Exposure to high concentrations of hydrogen peroxide on the order of 1 mM or strong pro-oxidative changes in the intracellular thiol/disulfide redox state will generally lead to increased tyrosine phosphorylation in numerous proteins (232,409, 491-493, 541). This effect is to some extent, albeit not exclusively, the consequence of the oxidative inhibition of protein tyrosine phosphatases. Massive inhibition associated with increased net phosphorylation of receptor tyrosine kinases is induced by various types of strong oxidative stress, including high doses of ROS, UV irradiation, or alkylating agents (43, 44, 139,217, 245, 251, 313,555). Protein tyrosine phosphatases counteract the effect of protein tyrosine kinases and reset membrane receptors after ligand-induced autophosphorylation. The EGF receptor, for example, is normally dephosphorylated at all tyrosine residues in <1 min after ligand-induced autophosphorylation (313), but this dephosphorylation is retarded by high concentrations of hydrogen peroxide on the order of 1 mM or other inducers of oxidative stress. A protein tyrosine phosphatase was also shown to regulate the activation of the EGF receptor (454).

All protein tyrosine phosphatases share a common sequence motif with a catalytically essential cysteine residue in the active center (42). Biochemical evidence suggests that inhibition of catalytic activity may proceed in either of two ways (Fig. 8). Hydrogen peroxide at concentrations on the order of 1 mM converts this cysteine residue into cysteine sulfenic acid (Cys-SOH) and thereby inactivates the enzyme, as demonstrated with three distinct protein tyrosine phosphatases (139). This effect is facilitated by vanadate and its derivative pervanadate. Alternatively, the redox-sensitive cysteine residue may be converted by glutathione disulfide into a mixed disulfide with concomitant loss of catalytic activity (43). Cysteine sulfenic acids are highly reactive and are expected to react with glutathione at its relatively high intracellular concentration. Therefore, it is reasonable to assume that ROS-induced oxidation will also lead to the rapid glutathionylation of the redox-sensitive cysteine moiety (Fig.9). Collectively, these findings indicate that protein tyrosine phosphatase is a typical example of a regulatory protein that responds either to changes in ROS concentrations or changes in the intracellular thiol/disulfide redox state, as was also found in the case of the bacterial OxyR protein (see Table 2). The cysteine-to-serine mutation at the catalytic site inactivates catalytic activity but not substrate binding activity of the phosphatases (182, 191, 285,556). The physiological relevance of the oxidative inhibition of the tyrosine phosphatases, however, is still controversial in view of the relatively high ROS concentrations that are typically required to inhibit the tyrosine phosphatases.

Fig. 9.

Fig. 9.Alternative pathways of protein tyrosine phosphatase inhibition by ROS or by changes in the thiol/disulfide redox state. A cysteine residue in the catalytic site of the phosphatase is critical for its catalytic activity. Inactivation can occur either by reaction with hydrogen peroxide to form a sulfenic acid derivative or by reaction with glutathione disulfide, resulting in glutathiolation of the critical cysteine residue. Reactivation may occur by reaction with reduced glutathione or other thiol compounds.


C.  Role of ROS in the Regulation of Insulin Receptor Kinase Activity

The most important insulin-responsive tissues are liver, skeletal muscle, and adipose tissue. In these tissues insulin controls several physiologically important functions, including the rate of glucose uptake, intracellular glucose metabolism, lipid metabolism, and the synthesis of proteins at the transcriptional and translational level (421).

Signaling by insulin requires autophosphorylation of the insulin receptor kinase at Tyr-1158, Tyr-1162, and Tyr-1163 (146,184, 266, 471, 612,616). In intact cells high concentrations of hydrogen peroxide on the order of 1 mM pervanadate and thiol-reactive agents induce insulin-like effects in the absence of insulin (121, 175, 240,247, 564, 620). Whenever tested, these effects were found to involve the insulin-independent tyrosine phosphorylation of the insulin receptor β-chain (175, 247, 564). In view of the marked inhibition of protein tyrosine phosphatases by hydrogen peroxide at concentrations on the order of 1 mM (see sect.v B), it is likely that activation of the insulin receptor by similarly high concentrations of hydrogen peroxide may be mediated, at least in part, by inhibition of tyrosine phosphatases (247).

Lower and physiologically relevant concentrations (<0.1 mM) of hydrogen peroxide are not sufficient to trigger the autophosphorylation of the insulin receptor in the absence of insulin, but do enhance the response to 100 nM insulin (496), indicating that the redox signal has a coregulatory function in insulin receptor activation under physiologically relevant conditions. This property is shared by other ROS-responsive signaling pathways. Because hydrogen peroxide production can be induced by insulin (321,380, 403), the redox effect appears again to be part of a positive feedback regulation (see Fig.10) that is reminiscent of the redox control of the EGF receptor. Krieger-Brauer et al. (321) reported that insulin-mediated stimulation of NADPH oxidase may not require the kinase activity of the insulin receptor. This finding remains to be confirmed. Regardless of this detail, the stimulation of ROS production through the insulin receptor and the redox sensitivity of insulin receptor kinase activity both suggest a certain degree of cooperativity between the insulin receptor and other membrane receptors, as illustrated in the case of the EGF receptor (Fig. 8).

Fig. 10.

Fig. 10.Role of ROS in insulin receptor kinase activation. Interaction of the insulin receptor with its ligand causes the activation of NAD(P)H oxidase and the production of superoxide and hydrogen peroxide. Hydrogen peroxide, in turn, is involved in the autophosphorylation and activation of the insulin receptor kinase. As in the case of the EGF receptor (see Fig. 8), this process is expected to benefit from activation of NAD(P)H oxidase by other types of membrane receptors in a cooperative way.

A combination of molecular modeling and functional studies indicated that the ROS-mediated enhancement of insulin receptor autophosphorylation involves cysteine residues within the receptor kinase domain itself (495, 496). The crystal structures of several protein kinase species, including the three times phosphorylated insulin receptor kinase domain (IRK-3P) (266), cAPK (656), Lck (643), c-Src (642, 622), Hck (523), and FGFRK (395), show structurally similar kinase domains with similar ATP binding sites. The crystal structure of the nonphosphorylated insulin receptor kinase domain (IRK-0P) differs from other kinase structures and exhibits an atypical position of the activation loop, which prevents productive ATP binding (267). However, autophosphorylation can be mediated through binding of phosphate donors to different binding sites in a process requiring the synergistic action of hydrogen peroxide. Three-dimensional models of the nonphosphorylated insulin receptor kinase domain revealed that conversion of any of the four cysteine residues 1056, 1138, 1234, and 1245 into sulfenic acid results in structural changes that render the well-known catalytic site at Asp-1132 and Tyr-1162 accessible for a phosphate donor from a direction different from that of the known ATP binding site. In addition, Tyr-1158 is brought into close contact with Asp-1083, suggesting that Tyr-1158 is using Asp-1083 rather than Asp-1132 as the catalytic amino acid for its autophosphorylation (495). Conversion of one of the four cysteine residues 1056, 1138, 1234, or 1245 into anS-nitroso derivative, i.e., the putative product after reaction with NO, leads to similar structural changes. Functional studies showed, however, that insulin-induced autophosphorylation of the insulin receptor β-chain is strongly inhibited by various NO donors (496). This effect can be tentatively explained by the assumption that S-nitrosylation prevents the further processing needed for activation of the insulin receptor domain. Taken together, these findings suggest that the apparent redundancy of functionally relevant cysteine residues in distinct regions of the insulin receptor protein may have resulted from an evolutionary process that increased the probability of a regulatory interaction with ROS.

As in the case of several other signaling processes (see Table 2), insulin responsiveness of insulin receptor kinase activity is amplified not only by exposure to ROS but also by a pro-oxidative shift in the intracellular glutathione redox state (494). Antioxidants such as butylated hydroxyanisol or the glutathione precursor N-acetylcysteine were found to inhibit the induction of insulin receptor kinase activity (494).

A decrease in intracellular glutathione levels resulting from treatment with buthionine sulfoximine has by itself little or no effect on insulin receptor activation but augments the effect of moderate concentrations of hydrogen peroxide (E. Schmid and W. Dröge, unpublished observations) or the stimulatory effect of the glutathione reductase inhibitor 1,3-bis(2-chloroethyl)-1-nitrosourea (BCNU) (494). Higher concentrations of hydrogen peroxide, however, tend to inhibit rather than enhance insulin receptor signaling and autophosphorylation (231; Schmid and Dröge, unpublished observations). With the use of glucose oxidase, it was also shown that prolonged exposure to moderate concentrations of hydrogen peroxide decreases insulin receptor signaling (576).

D.  Activation of Cytoplasmic Protein Kinases by ROS

In line with the strong inhibition of protein tyrosine phosphatases by hydrogen peroxide at high concentrations on the order of 1 mM, several authors reported a massive increase in tyrosine phosphorylation of various cellular proteins under similar highly oxidative conditions (232, 409,491-493). The activity of several protein tyrosine kinases such as Lck, Fyn, Syk, and ZAP70 is enhanced under such conditions (75, 239, 409,499-501). The physiological relevance of these strongly oxidizing conditions is still unclear. At concentrations of <0.2 mM hydrogen peroxide or after moderate sulfhydryl oxidation, the increase in phosphorylation is restricted to one or a few prominent proteins (409). Exposure of T cells to 0.15 or 0.5 mM hydrogen peroxide yields a single major phosphorylated protein, the Src family protein tyrosine kinase p56lck. This phosphorylation is associated with the induction of p56lck kinase activity (409). Analogous to this finding, hydrogen peroxide was found to stimulate JAK2, a critical mediator for the activation of the Ras/Raf/ERK pathway, in normal murine fibroblasts, in fibroblasts derived from transgenic mice deficient in p60src, but not in fibroblasts deficient in p59fyn (3). Moreover, Ras activation by hydrogen peroxide was found to be impaired in p59fyn-deficient cells but not in p60src-deficient cells. Studies on a human T cell line showed that the MAPK species c-Jun NH2-terminal kinase (JNK) and p38 are activated by mild intracellular thiol oxidation in a tyrosine kinase-dependent manner (248). It was also demonstrated that the kinase activity of p59fyn is activated by low micromolar concentrations of glutathione disulfide and that the Src family kinases p59fyn and p56lckare strongly activated by a mild oxidative shift of the intracellular thiol/disulfide redox state (248). Analogous findings were obtained in studies of rat vascular smooth muscle cells and fibroblasts where the protein tyrosine kinase Src was found to be critically involved in JNK activation by 0.03 or 0.1 mM hydrogen peroxide (651). The molecular details of the activation of Src family kinases by mildly oxidizing conditions, however, are not clearly understood.

E.  Oxidative Activation of MAPK Cascades

MAPK signaling cascades are regulated by phosphorylation and dephosphorylation on serine and/or threonine residues and respond to activation of receptor tyrosine kinases, protein tyrosine kinases, receptors of cytokines and growth factors, and heterotrimeric G protein-coupled receptors. Numerous studies with various experimental systems show that in particular the MAPK species JNK and p38 are strongly activated by ROS or by a mild oxidative shift of the intracellular thiol/disulfide redox state (2,20, 222, 248, 355). The extracellular signal-regulated kinase 1 (ERK-1) and ERK-2 were found to be activated in vascular smooth muscle cells by superoxide but not by hydrogen peroxide (35). Angiotensin II induces the induction of superoxide and hydrogen peroxide and activates ERK-1, ERK-2, and p38 MAPK (2, 590). PDGF was found to induce the activation of ERK-1 and ERK-2 (557). JNK and p38 MAPK were shown to be activated by hydrogen peroxide in perfused rat hearts (112). As described in sectionv D, the redox sensitivity of JNK and p38 MAPK is mediated at least to some extent by the oxidative activation of upstream tyrosine kinases of the Src family (452).

In addition, the JNK/glutathione S-transferase Pi (GSTp) complex has also been identified as a redox-responsive signaling element. In normal growing cells of the mouse fibroblast cell line 3T3–4A, JNK is associated with and catalytically inhibited by glutathione-S-transferase (GSTp) (12). Complex formation between GSTp and JNK limits the degree of Jun phosphorylation under nonstressed conditions. Exposure to low micromolar hydrogen peroxide concentrations causes the oligomerization of GSTp and the dissociation of the GSTp-JNK complex, indicating that JNK inhibition requires monomeric GSTp.

The apoptosis signaling-regulating kinase 1 (ASK1) plays a role in the activation of MKK3/6, MKK4/MKK7, and the MAPK species p38 and JNK (270). This leads ultimately to the phosphorylation of ATF2, c-Jun, and p53 (11, 189,275). Screening for ASK1-associated proteins has led to the identification of Trx as the redox-sensitive target molecule (481). Under normal conditions, Trx binds to the NH2-terminal domain of ASK1 and inhibits its kinase activity. Deletion of the Trx-binding NH2-terminal residues of ASK1 renders it constitutively active and no longer responsive to the inhibitory effect of Trx. ROS induce the dimerization of Trx and its dissociation from ASK1, followed by multimerization of ASK1 and activation of its kinase activity (205,353).

F.  Oxidative Activation of Protein Kinase C Isoforms

The serine/threonine kinase protein kinase C (PKC)-α is involved in signal transduction to various effector pathways that regulate transcription and cell cycle control. Certain PKC isoforms, including PKC-α are typically activated by the lipid second messenger diacylglycerol (416, 430). This effect is mimicked by phorbol esters (57). The binding site for diacylglycerol or phorbol esters is located in a conserved cysteine-rich region within the NH2-terminal C1 domain (83, 430, 455,457).

Alternatively, PKC-α and certain other PKC isoforms can be activated by hydrogen peroxide in a phospholipid-independent process that involves tyrosine phosphorylation in the catalytic domain (204, 316). The oxidative activation of PKC-α is significantly enhanced by vitamin A and certain derivatives such as 14-hydroxy-retroretinol which bind to a zinc finger domain in the conserved cysteine-rich region of the regulatory domain (263). The serine/threonine kinase cRaf contains an analogous cysteine-rich domain that binds phosphatidylserine instead of diacylgylcerol (185). Like PKC-α, cRaf is also activated by ROS in a retinol-dependent fashion (263, 274). Vitamin A or its derivatives 14-hydroxy-retroretinol and 13,14-dihydroxyretinol are strictly required for lymphoid cell growth and viability (78,141, 193, 627).

α-Tocopherol (vitamin E), in contrast, was found to inhibit the translocation of PKC into the membrane and the phosphosrylation of its 80-kDa protein substrate in smooth muscle cells (34,62). This effect may be involved in the α-tocopherol-mediated inhibition of vascular smooth mucle cell proliferation (62) and of NF-κB activation in Jurkat T cells (561). RRR-β-tocopherol, in contrast, does not inhibit PKC activation but prevents the effect of RRR-α-tocopherol (34).

G.  ROS-Induced Changes in Cytosolic Ca2+Concentrations

Changes in the cytosolic Ca2+ level play a role in the modulation of several intracellular signal pathways, including PKC-α and calmodulin-dependent signal pathways (111). These pathways have also been implicated in apoptotic processes. The cytosolic Ca2+ level can be increased by ROS in various cell types through the mobilization of intracellular Ca2+stores and/or through the influx of extracellular Ca2+(150, 155, 226,327, 427-429, 476). The ROS-mediated increase in the cytosolic Ca2+concentration contributes to the oxidative stress-mediated activation of PKC-α (340) and to the transcriptional induction of the AP-1 proteins c-Fos and c-Jun (118, 367, 520).

H.  Activation of the Transcription Factor AP-1

The transcription factor AP-1 is typically composed of c-Fos and c-Jun proteins and is implicated in differentiation processes. In T lymphocytes AP-1 is involved in the expression of the interleukin-2 gene and other immunologically relevant genes (372, 552). Many different oxidative stress-inducing stimuli, e.g., relatively low concentrations of hydrogen peroxide, UV light, γ-irradiation, and interleukin-1, lead to AP-1 activation by different mechanisms (23,144, 389, 419,548).

The mRNAs of c-Fos and c-Jun are induced by relatively small amounts of hydrogen peroxide, superoxide, NO, and other inducers of oxidative stress (117, 283, 389,401, 402, 419,520). If Jurkat T lymphocytes are treated with 200 μM hydrogen peroxide in the absence of a mitogenic stimulus, expression of c-Jun and transient expression of c-Fos are enhanced without subsequent induction of interleukin-2 production (52). In contrast, treatment of the murine T cell line ESb with 200 μM hydrogen peroxide causes an increase in AP-1 transcription factor activity and the expression of c-Jun and interleukin-2 (358). The oxidative activation of AP-1 transcription factor activity is based on oxidative activation of JNK. This MAPK phosphorylates serine residues 63 and 73 of the NH2-terminal transactivation domain of c-Jun, a domain which is required for functional activation (295,297).

As for the redox-sensitive bacterial OxyR protein (see sect.iii B2) and activation of the human insulin receptor kinase (see sect. v C), the activation of AP-1 and its upstream signaling cascades can also be enhanced by a change in the intracellular glutathione redox state (192,248). A mild oxidative shift of the redox state by three different oxidants resulted in a strong enhancement of JNK and p38 MAPK activity but not so for ERK-1 and ERK-2 (248). JNK and p38 MAPK are also activated by UVA irradiation and by singlet oxygen (311, 312). High concentrations of hydrogen peroxide (∼1 mM) also result in the activation of ERK-1 and ERK-2 in Jurkat T cells and in cardiac myocytes (14,213). The differential activation of JNK, p38 MAPK, and ERK by various types of ROS, or RNS or by UV irradiation in different cell types has been reviewed by Klotz et al. (312).

I.  Activation of the Transcription Factor NF-κB

NF-κB is involved in the induced expression of the interleukin-2 gene and in a wide variety of biological responses. In particular, it is implicated in inflammatory reactions, growth control, and apoptosis (38, 39) and is the first eukaryotic transcription factor shown to respond directly to oxidative stress in certain types of cells (501). NF-κB activation was shown to be inhibited by antioxidants such as cysteine (389,391, 490, 505,542). In various cell types such as the Würzburg subclone of T cells, L6 skeletal muscle myotubes, human breast MCF-7 cells, and 70 Z/3 pre-B cells, NF-κB is activated by micromolar concentrations of hydrogen peroxide (368,391, 501, 502,512). A moderate pro-oxidative shift in the glutathione redox state was also found to enhance NF-κB activation in Molt-4 and Jurkat T cell lines (192, 248). However, NF-κB activation is not strictly dependent on ROS. Several signal cascades have recently been identified that can activate NF-κB without involvement of ROS, and hydrogen peroxide does not induce NF-κB activation in all cell types (65, 66,350, 560). In Molt-4 cells, NF-κB activity was found to be enhanced by 30 μM hydrogen peroxide if cultured in NCTC 135 medium, i.e., a culture medium with relatively low concentrations of free amino acids, but not in RPMI 1640 medium (160).

Taken together, the large body of experimental evidence now suggests that ROS comodulates the activation of NF-κB in certain cell types under certain conditions. At least two mechanisms contribute to this effect. The first one involves the ROS-mediated enhancement of IκB degradation. Treatment of HeLa cells with hydrogen peroxide induces the appearance of a modified form of the inhibitor IκBα which is rapidly degraded upon appropriate stimulation unless proteolysis is inhibited by a proteasome inhibitor (319). Treatment with antioxidants was found to block IκBα degradation after stimulation with TNF, TPA, or lipopolysaccharide (53, 320, 368,580). Also, overexpression of glutathione peroxidase was found to abolish the TNF-mediated accumulation of the modified form of IκBα (320), and treatment with the antioxidantN-acetylcysteine was shown to inhibit NF-κB activation and IκBα degradation but not IκBα phosphorylation (350). In view of the role of ROS in the regulation of proteosome-mediated protein catabolism (see sect.ii C5), it is reasonable to assume that the illusive mechanism of the redox control of NF-κB is mediated at least partly by the more general mechanism of ROS-mediated enhancement of proteolysis according to the model in Figure 2. A second mechanism involves the oxidative enhancement of upstream signal cascades. Increased phosphorylation of IκBα was found in EL4 T cells after exposure to 300 μM hydrogen peroxide (499), and a similar induction of IκBα phosphorylation and activation of the IκB kinase IKKα was seen in Jurkat T cells after a moderate pro-oxidative shift in the intracellular glutathione redox state (248).

J.  Importance of the Intracellular Glutathione Level

In all cases examined, the redox regulation of membrane receptor-dependent signals was found to be strongly dependent on the intracellular glutathione level. The enhancement of insulin receptor kinase activation by inhibition of glutathione biosynthesis (see sect. v C) is an example. Similarly, interleukin-2 production by resting small T cells of the mouse was found to be enhanced by superoxide or hydrogen peroxide only if the cells were exposed simultaneously to physiologically relevant millimolar concentrations of l-lactate (472), a macrophage product (33) known to decrease intracellular glutathione levels in lymphocytes (473, 475). Because the lactate derivative pyruvate is an effective scavenger of the glutathione precursor cysteine (159, 286) and because the addition of glutathione abolishes the lactate effect (475), it is reasonable to assume that the lactate-dependent enhancement of interleukin-2 production is due to glutathione depletion.

The biosynthesis of glutathione is mainly determined by the concentrations of the precursor amino acids (465) and competes with protein synthesis for the available amino acids. A redox mechanism that links receptor-mediated signals to low intracellular glutathione levels favors therefore those cells that are engaged in highly active protein synthesis.

K.  Differential Redox Requirements in the Induction and Execution of Signal Cascades

1.  Signal amplification by a late shift to reducing conditions

Whereas the induction of signaling cascades leading to the activation of the transcription factors NF-κB and AP-1 is typically enhanced by pro-oxidative conditions, the final execution of the signaling process requires relatively reducing conditions. To induce the expression of specific genes, the transcription factors must bind to specific sequences of highly acidic, negatively charged DNA. Therefore, the binding region of the transcription factors typically contains an accumulation of positively charged amino acids (reviewed in Ref. 160). This accumulation of positive charges can stabilize deprotonated thiol groups, making any cysteine group in the DNA-binding region highly susceptible to oxidation. For example, the transcription factor NF-κB and other members of the NF-κB family share a characteristic sequence motif with one cysteine and three arginine residues in the DNA-binding region (198, 302, 326,374). The DNA-binding activity of NF-κB is, therefore, easily inhibited in vitro by oxidation and reactivated by thiols such as 2-mercaptoethanol, dithiothreitol, and thioredoxin (Trx) (192, 577). Physiologically relevant concentrations of glutathione disulfide inhibit DNA binding of NF-κB, even in the presence of a severalfold excess of thiols, and DNA-binding activity was found to be restored by physiologically relevant concentrations of reduced Trx (192). NMR studies have demonstrated the direct binding of a Trx peptide to the DNA-binding domain of NF-κB (453).

Similarly, oxidation of a conserved cysteine residue in the DNA-binding region adjacent to the leucine zipper in AP-1 abolishes its binding to DNA (637). The DNA-binding activity of AP-1 was found to be restored by Trx and the nuclear signaling protein redox factor-1 (Ref-1) (1, 637,638) and enhanced by transient overexpression of Trx in vivo (490). Compared with NF-κB, AP-1 is less sensitive to oxidative inhibition by glutathione disulfide in vitro but more sensitive to oxidized Trx (192). Thus, in intact cells, the elevation of intracellular oxidized glutathione (GSSG) levels by inhibition of glutathione reductase inhibits selectively the DNA-binding activity of NF-κB and not that of AP-1 (192).

To reconcile these seemingly conflicting redox requirements for the induction and execution phases of the signaling process, a delicately balanced intermediate redox state is needed (160). The oxidative enhancement of interleukin-2 production in macrophage-depleted T-cell populations is a case in point. Interleukin-2 production is strongly enhanced by 10 μM hydrogen peroxide but completely inhibited by higher concentrations of hydrogen peroxide (100 μM) (52, 436,472). Analogous to the inhibition of insulin receptor signaling by excessive or prolonged ROS exposure, prolonged exposure to weak oxidative stress also suppresses anti-CD3-induced interleukin-2 production, protein tyrosine phosphorylation, and other parameters of cell activation (180). Moreover, the induction of NF-κB in Molt-4 cells is enhanced by moderate concentrations of the glutathione reductase inhibitor BCNU but inhibited at higher concentrations (192). A study on healthy human subjects showed that persons with intermediate intracellular glutathione levels (20–30 nM/mg protein) had higher CD4+ and CD8+T-cell numbers than persons with either lower or higher glutathione levels (307).

One of several mechanisms that satisfies the different redox requirements for induction and execution of redox-sensitive signal cascades is based on the redox-mediated nuclear translocation of Trx. Ionizing radiation, which causes oxidative stress, was found to increase AP-1 DNA-binding activity in Jurkat T cells and HeLa cells by increasing the nuclear level of Trx.

In addition, lymphocytes are extremely mobile, and their migration from an oxidative (i.e., inflammatory) into a more reducing environment is expected to help the cells to meet differential redox requirements during the induction and execution phases and to enhance thereby the immunological response. Experiments with cultured T cells showed that activation of AP-1 and NF-κB transcription factors can be markedly enhanced if the cells are stimulated under relatively oxidative conditions and shifted after 1 h to more reducing conditions for optimal DNA binding (192). In the in vivo setting lymphocytes experience such a shift whenever they move out of the inflammatory environment into more reducing tissues as illustrated schematically in Figure 11. This requirement for a shift in the redox condition may be viewed as an additional fail-safe mechanism against inappropriate activation of (auto)immune reactions.

Fig. 11.

Fig. 11.Distinct redox requirements in the induction and execution of signal cascades: hypothetical model. The antigen-responsive signaling cascades of lyphocytes are strongly enhanced by the ROS produced in large quantities by macrophages in the inflammatory environment. Glycolytically active macrophages produce also large amounts of lactate, which, in turn, induces a decrease in the intracellular glutathione level of the lymphocyte. Experiments with Molt-4 cells have shown that the 4-O-tetradecanoylphorbol 13-acetate (TPA)-induced activation of activator protein 1 (AP-1) and nuclear factor κB (NF-κB) DNA-binding activity is inhibited if the cells are subjected to reducing conditions (1 mM dithiothreitol) before TPA treatment but strongly enhanced if the cells are subjected to reducing conditions 1 h after TPA stimulation. Moreover, the proliferative response and the induction of many types of immunological responses are strongly dependent on relatively high intracellular glutathione levels. The delivery of reduced cysteine by certain types of macrophages to the lymphocytes has been shown to play an important role in the maintenance of adequate glutathione levels in lymphocytes.

Lymphocytes require in the execution phase and for the subsequent DNA synthesis response large amounts of the glutathione precursor cysteine. Even a moderate glutathione deficiency was found to impair various immunological functions (reviewed in Ref. 160). Because lymphocytes have only weak transport activity for cystine, they take advantage of certain types of macrophages that are capable of taking up cystine and releasing large amounts of cysteine into the extracellular space (199). The importance of this process is illustrated by the fact that lamina propria macrophages lack the ability to release cysteine, thus ensuring the physiological unresponsiveness of the lamina propria T lymphocytes to antigen exposure in the intestinal microenvironment (524). The hyperresponsiveness of lamina propria T lymphocytes in patients with inflammatory bowel disease is associated with the invasion of blood monocytes into the inflamed area.

In the execution phase, pyruvate can serve also as a major endogenous scavenger for hydrogen peroxide (151,159). Activated and cycling T cells cover a large proportion of their energy demand by glycolytic metabolism, which generates substantial amounts of pyruvate (68,209, 609).

2.  Induction of the cell cycle inhibitor p21 by ROS

The cell cycle inhibitor p21 (WAF1, CIP1, or sdi1) plays a central role in the control of the cell cycle by interacting with multiple targets including cyclin-dependent kinases (236,640). ROS were shown to induce p21 gene expression by an unknown mechanism that may involve p53 (123,369, 422). This underscores the importance of ROS scavenging mechanisms in the proliferative phase of cellular responses.

L.  Key Messages From Section v

Signal transduction from various membrane receptors is enhanced by ROS.

Membrane receptors of various growth factors, cytokines, or other ligands induce positive feedback effects on signal transduction from these receptors by triggering concomitantly the activation of NAD(P)H oxidases.

Enhancement of signal transduction from a given receptor by stimulation of ROS production through this or other receptors provides the basis for cooperativity. Because hydrogen peroxide has a relatively long half-life and can cross membranes, this cooperativity may even extend to other cells in the vicinity. In addition, the membrane receptor may function simultaneously as a sensor for extracellular signals and as a sensor for the inner metabolic state of the individual cell.

Oxidative enhancement of membrane receptor signaling and corresponding downstream signaling pathways are not well characterized at the molecular level but are likely to involve simultaneously the oxidative modification of several different redox-sensitive signaling proteins.

Certain signaling cascades involving protein tyrosine kinases can be enhanced by oxidative inhibition of protein tyrosine phosphatases. The molecular basis of this effect is relatively well-defined and involves the oxidative derivatization of a catalytically essential cysteine residue in the enzyme's active center. Therefore, tyrosine phosphatases can also be inactivated by changes in the intracellular thiol/disulfide redox state. Because most studies have been performed with relatively high concentrations of hydrogen peroxide (∼1 mM) or strong pro-oxidative changes in the intracellular thiol/disulfide redox state, the physiological relevance of this interesting mechanism needs further clarification.

Stimulation of the insulin receptor tyrosine kinase activity by insulin can be enhanced by relatively moderate (micromolar) concentrations of hydrogen peroxide or mild pro-oxidative changes in the intracellular thiol/disulfide redox state. Molecular modeling studies suggest that this redox effect may be mediated by the oxidative derivatization of any of four cysteine residues in the tyrosine kinase domain of this membrane receptor. These findings need to be confirmed by crystallographic analysis or other experimental techniques.

Signaling pathways involving JNK, p38 MAPK, and the transcription factor AP-1 are strongly responsive to redox regulation. The oxidative enhancement of Src-family tyrosine kinases and certain PKC isoforms as well as the dissociation of the JNK/GSTp and ASK1/Trx complexes may together contribute to the redox sensitivity of these pathways.

Activation of NF-κB is another well-studied model of redox regulation. ROS or changes in the thiol/disulfide redox state are not strictly required for NF-κB activation but induce or amplify NF-κB activation in various cell types under various conditions. At least two different mechanisms contribute to the enhancement of NF-κB activation. One of these mechanisms is based on the enhanced proteolytic degradation of the NF-κB inhibitor IκB after exposure to ROS. The second mechanism involves the increase in IκB kinase-α activity after exposure to hydrogen peroxide or to pro-oxidative changes in the intracellular glutathione redox state.

The in vivo relevance of redox-sensitive signaling cascades is strongly suggested by the large number of NAD(P)H oxidase isoforms (see sect. iv A) and by the dysregulation of physiological responses in various disease-related oxidative stress conditions (see sect. vi). However, the relative contribution of individual redox-sensitive signaling proteins to redox-regulated processes in vivo is presently obscure.

The DNA-binding activity of many if not most transcription factors is sensitive to oxidative conditions. Physiological responses require, therefore, a delicate balance between the pro-oxidative conditions that are needed to strengthen the signaling cascades and the reducing conditions that are needed for the execution of these signals.

VI. ROLE OF REACTIVE OXYGEN SPECIES IN SENESCENCE, STRESS CONDITIONS, AND DISEASE: PATHOPHYSIOLOGICAL IMPLICATIONS OF REDOX REGULATION

A.  Mediators of Excessive ROS Production

Redox-regulated physiological processes are inevitably sensitive against excessive ROS production by any source. Such excessive levels of ROS may be generated either by excessive stimulation of the otherwise tightly regulated NAD(P)H oxidases (see sect.iv A) or by other mechanisms that produce ROS "accidentally" in a nonregulated fashion. These latter mechanisms include the production of ROS by the mitochondrial ETC or by xanthine oxidase.

1.  The mitochondrial electron transport chain as a source of ROS

The mitochondrial electron transport chain (ETC) is a relatively well-investigated source of ROS. A major site for the univalent reduction of molecular oxygen to superoxide is ubisemiquinone, a component of the ETC in the mitochondrial matrix (63,64, 86, 101, 122,322, 360, 361, 417,584). Practically all cells and tissues convert continuously a small proportion of oxygen into superoxide by this mechanism. If carefully deprived of contaminating superoxide dismutase, submitochondrial particles generate superoxide at a rate of 4–7 nmol·min−1·mg protein−1, suggesting that the mitochondrial membrane is the quantitatively most important physiological source of superoxide in higher organisms (101). With acetone-extracted submitochondrial particles it was found that supplementation with exogenous ubiquinones led to ROS production rates that were linearly related to the amount of reducible quinone (63, 64). NADH-ubiquinone reductase and ubiquinol-cytochrome creductase, which contain ubisemiquinone as an important constitutent, were shown to generate superoxide and hydrogen peroxide (86). Mitochondrial ROS production has been implicated also in TNF-mediated oxidative stress (504,505).

2.  ROS production by xanthine oxidase

Xanthine oxidase generates superoxide by converting hypoxanthine into xanthine and xanthine into uric acid. The enzyme is derived from xanthine dehydrogenase by proteolytic cleavage. Under normal conditions, xanthine oxidase accounts for only a minor proportion of total ROS production (reviewed in Ref. 101). However, ROS production by xanthine oxidase has been observed in TNF-treated endothelial cells (188) and has also been implicated as a major source of oxidative stress under certain disease conditions such as ischemia and reperfusion (see sect. vi O).

3.  Dopamine as a source of ROS in the central nervous system

ROS generated by oxidation of dopamine has been implicated in the aging-related destruction of dopaminergic neurons and especially in Parkinson's disease (reviewed by Luo and Roth, Ref. 365).

4.  Other sources of ROS

In addition to the mechanisms described here, there are various other enzymic and nonenzymic mechanisms of ROS production as reviewed elsewhere (101, 122, 365,526).

B.  The Free Radical Theory of Aging

Multicellular organisms generally undergo qualitative changes with time (aging) that are associated with progressive degeneration of biological functions, increased susceptibility to diseases, and increased probability of death within a given time period. The widely popular free radical theory of aging (234) states that the age-related degenerative process is to a large extent the consequence of free radical damage. Genetic evidence linking oxidative stress to life span has been obtained for different animal species. InCaenorhabditis elegans, the daf-2 mutation causes longevity by increasing Mn-SOD expression (260). Catalase is required to extend the life span in daf-C andclk-1 mutants of C. elegans (566), and synthetic compounds exhibiting both SOD- and catalase-like activities enhance the mean life span of wild-type worms by 44% (388). Moreover, a mev-1 (kn1)/cyt-1 mutation, which inactivates succinate dehydrogenase cytochrome b, was found to render C. elegans susceptible to oxidative stress and, as a result, leads to premature aging (277). Themth mutant of Drosophila has a significantly extended life span and increased resistance to a free radical generator (351). An extended life span was also observed inDrosophila strains with extra copies of genes encoding SOD and catalase (431, 439). Last but not least, mice carrying a mutation in the p66shc protein were found to have an increased life span associated with increased resistance to oxidative stress (390). It is not within the scope of this review to discuss radical-mediated oxidative damage, but the following points ought to be addressed: 1) What is the evidence that old age and/or disease conditions are associated with changes in ROS levels? 2) What are the consequences of such changes with respect to redox-regulated processes? 3) What are the causes for changes in ROS generation? 4) What are the chances that therapeutic strategies can ameliorate or reverse the increase in ROS generation?

C.  Indications for an Age-Related Increase in ROS Levels

Although ROS production is difficult to measure in biological tissues, there are various indirect manifestations of oxidative stress in old age, including lipid peroxidation, DNA oxidation, protein oxidation, and a shift in the redox states of thiol/disulfide redox couples such as glutathione, cysteine, and albumin (22,49, 225, 343, 344,418, 463, 537-539,654). The finding that the skeletal muscle tissue of rhesus macaques shows massive age-related manifestations of oxidative damage (654) is of special significance, because the loss of skeletal muscle mass (sarcopenia) is one of the hallmarks of age-related wasting and a major cause of psychological stress, financial burden, and loss of social functions in elderly human subjects (72, 93, 332).

An increasing body of evidence suggests that 1) the mitochondrial genome may be particularly susceptible to oxidative damage during aging and 2) mitochondrial DNA deletion mutations may contribute to the fiber atrophy that causes sarcopenia. In a study on rectus phemoris muscle, fibers from 5- to 38-mo-old rats revealed that muscle fibers harboring mitochondrial deletions often display atrophy and increased steady-state levels of oxidative nucleic damage (605).

The published data do not allow us to distinguish whether these age-related changes result from an age-related accumulation of oxidative damage or from an age-related increase in ROS production per unit time. However, age-related changes corresponding to the major homeostatic control mechanisms illustrated in Figure 2 have indeed been observed and may provide an indirect indication for an age-related increase in ROS production. The analysis of gene expression profiles by oligonucleotide arrays representing 6,347 genes revealed aging-related changes indicative of a stress response in skeletal muscle and brain tissue of mice (343,344). All three tissues tested, i.e., skeletal muscle, neocortex, and cerebellum, showed in old age an increased expression of heat shock factors and other oxidative stress-inducible transcripts. The two brain tissues showed, in addition, increased expression of immunologically relevant transcripts indicative of an inflammatory response (344). Another study examined skeletal muscle tissue from rats and revealed a significant age-related increase in the hydrogen peroxide scavenging enzymes catalase and glutathione peroxidase (284,596). A conspicuous pro-oxidative shift in the plasma thiol/disulfide redox state has been observed in human subjects between the 3rd and the 10th decade of life (224). Nevertheless, the aging process in the elderly is remarkably slow, indicating that even the elderly have adopted a nearly stable state of redox homeostasis.

A shift in the systemic thiol/disulfide redox state may have systemic consequences because several of the redox-sensitive signal cascades respond not only to direct exposure to ROS but also to changes in the thiol redox state (see Table 2). Therefore, an age-related increase in ROS generation at a given anatomical site may cause a redox-mediated dysregulation at multiple and distant anatomical sites of the organism. The plasma thiol/disulfide ratio may be an easy target for therapeutic intervention by oral N-acetylcysteine or other thiol compounds (225).N-acetylcysteine was previously shown to protect against age-related decreases in oxidative phosphorylation in liver mitochondria (393). A shift in the redox state may cause, among other consequences, changes in the sensitivity of oxygen sensors (W. Hildebrandt and W. Dröge, unpublished observations), a decrease in the plasma albumin level (225), and, last but not least, changes in cellular functions that are under the control of redox-sensitive signaling cascades. The widely investigated phenomenon of replicative senescence in fibroblasts may be a case in point.

It is necessary, however, to add a note of caution. Although it may be a plausible and attractive paradigm to postulate that aging-related changes in the thiol/disulfide redox state and the changes in gene expression profiles are the direct consequence of an age-related increase in the rate of ROS production, this cause-and-effect relationship remains to be proven.

D.  Replicative Senescence as a Putative Consequence of Redox-Mediated Dysregulation

Somatic cells divide in cell culture only for a finite number of generations. The eventual arrest of cell division is termed "replicative senescence" (241-243, 258,508, 625). Because the number of divisions of human dermal fibroblasts in vitro decreases with the age of the donor, cellular senescence is commonly seen as an in vitro correlate of the aging process (241). Replicative senescence of human fibroblasts in vitro is associated with morphological changes indicative of differentiation (48, 467). Studies with a histochemical marker suggested that senescent cells may accumulate with age in vivo (147). As in the case of fibroblasts, T cell lines and clones from older individuals typically show a smaller number of population doublings in vitro than observed for cells from younger persons (218, 384,604).

More recent studies suggested, however, that the replicative life span in vitro is not an inherent property of somatic cells but determined, at least in part, by the redox state of the microenvironment. A senescence-like growth arrest is rapidly induced by hydrogen peroxide (104), and similar effects were found for primary human diploid fibroblasts after a p21ras-mediated increase in intracellular ROS levels (342) and in fibroblast cultures exposed to 8-methoxypsoralen followed by UVA irradiation (252). Earlier studies with human diploid fibroblasts have shown that cells exhibit a prolonged life span if grown at low oxygen tension (435). Furthermore, experiments with buthionine-sulfoximine, a specific inhibitor of glutathione biosynthesis, showed that the intracellular redox state modulates the balance between self-renewal and differentiation in dividing glial precursor cells (536).

The evidence discussed above strongly suggests that the observed decrease in replicative capacity in vivo may be a result of the age-related increase in ROS levels and/or the progressive shift in the systemic thiol/disulfide redox state (see sect.vi C). This conclusion is supported by the observation that replicative senescence of human fibroblasts is associated with a dysregulation of the AP-1 transcription factor and with changes in the posttranslational modification of the Fos protein (464). Replication-incompetent cell types express new sets of genes, but the corresponding gene products do not appear to be functionally relevant. Regardless of whether these changes justify a comparison with a differentiation process or not, the observed pro-oxidative changes in elderly subjects (see sect.vi C) strongly suggest that aging of an organism may be associated with changes in redox-sensitive signaling pathways. These changes may well account for differentiation-like processes and a loss in replicative capacity. An additional complication may arise from the fact that senescent fibroblasts suffer from a marked inhibition of proteasome activity (530-532).

The conspicuous discrepancy between the cellular senescence of fibroblasts and the uncontrolled growth of malignant cells (see sect.vi I) may be explained by the presence or absence of the tumor suppressor PML, which is strictly required for cellular senescence (442). Defective functioning of PML was shown to be involved in the development of malignancies (61,143, 610).

E.  Oxidative Induction of Telomere Shortening

The discovery that telomeres get progressively shorter in aging human fibroblasts (233) has led to the popular hypothesis that telomere shortening may also be a major cause of cellular senescence. The expression of a telomerase transgene in cell lines derived from patients with Werner syndrome results in lengthened telomeres and replicative immortalization, thus indicating that the shortening of telomeres is a trigger of premature senescence in these cells (107). Several more recent reports showed that telomere shortening can be induced in fibroblasts by mild oxidative stress (533, 602, 603).

F.  Factors Contributing to Changes in ROS Production

1.  Age-related changes in mitochondrial complex IV activity and the effect of substrate availability on mitochondrial ROS production

The probability that molecular oxygen is reduced to superoxide rather than water is increased if the proton gradient at the mitochondrial matrix is high (see Fig.12) and the proper flux of electrons through the ETC energetically less favored. Because the proton gradient is coupled to the conversion of ADP into ATP (Fig. 12), mitochondrial ROS generation is particularly strong if the availability of ADP is low (64, 102, 362). Under these conditions the components of the ETC are largely in a reduced state. Because ATP consumption and APD availability are particularly low during periods of sleep when the muscular activity is low, the mitochondrial oxidative stress may be particulary high at night. Moreover, a study on the gastrocnemius muscle in mice revealed that the activity of complex IV decreases by 90% from 10 to 26 mo of age (142). This process is likely to impair significantly the proper flux of electrons through the ETC and to enhance thereby the mitochondrial production of superoxide.

Fig. 12.

Fig. 12.Effect of substrate availability on ROS production. Substantial amounts of superoxide are produced at the semiubiquinone component of the mitochondrial electron transport chain. The probability of ROS production is increased if the influx of electrons is high and the consumption of electrons by cytochrome oxidase in the mitochondrial complex IV relatively low. Consumption of electrons is typically low if the proton gradient is high, i.e., if only a few protons are being consumed by the ATP-generating system as a result of relatively high ATP and low ADP concentrations. This condition normally inhibits the glycolytic pathway and the influx of further energy substrates into the mitochondria. At high glucose concentrations, e.g., as a result of inadequate glycogen synthase (GS) activity, this control may be overridden, resulting in increased ROS production. PFK, phosphofructokinase.

The influx of electrons into the ETC, in turn, is determined by the availability of the electron donors NADH (complex I) and succinate (complex II). Submitochondrial particles depleted of succinate dehydrogenase are still capable of ROS production (63,360, 361), suggesting that the succinate dehydrogenase flavoprotein does not play an important role in ROS production.

The availability of NADH is determined by the availability of mitochondrial energy substrates such as acetyl CoA. An excess of substrates is normally prevented by the tight regulation of key enzymes in the glycolytic pathway (Fig. 12). Phosphofructokinase (PFK), one of the early enzymes in this pathway, is inhibited by ATP and citrate and couples thereby the generation of pyruvate and acetyl CoA to the cellular energy demand. The negative control of the glycolytic pathway can be overridden, however, by excess fructose-6-phosphate, the PFK substrate derived from glucose via glucose-6-phosphate. A disproportionate amount of the excess glucose taken up by muscle tissue in patients with non-insulin-dependent (type 2) diabetes mellitus in the postabsorptive state due to plasma hyperglycemia is converted glycolytically into lactate (reviewed in Ref. 131). In aorta endothelial cells, a pathologically relevant elevation of the extracellular glucose concentration was shown to cause a massive increase in ROS production by the mitochondrial ETC (415).

TNF-mediated wasting may be mediated, at least to some extent, by a TNF-induced increase in glycolytic activity and lactate production (46). Under certain conditions, NO and peroxynitrate may contribute to mitochondrial superoxide production by enhancing the build-up of semi-ubiquinone (87).

2.  Effect of caloric restriction on ROS production, age-related diseases, and life span

Because the rate of mitochondrial ROS production is significantly influenced by the availability of mitochondrial energy substrates, it is not surprising that dietary restriction is today the best investigated and most promising experimental strategy to increase life span and to improve the quality of life in old age (181,343, 344, 537, 539,613, 614, 654). Studies in several animal species have shown that caloric restriction ameliorates certain manifestations of oxidative stress and causes a substantial increase in life span. Dietary restriction in rodents also delays the onset of age-associated diseases, decreases the incidence of malignancies, and ameliorates the decline in mitochondrial complex IV activity (142, 613). The life-extending benefits of caloric restriction depended in all the cited studies on the prevention of malnutrition and a reduction in total caloric intake rather than in any particular nutrient (613).

3.  Can endurance exercise substitute for caloric restriction?

Whereas rigorous caloric restriction may be an unattractive regimen for human subjects, endurance exercise may yield similar effects with lower risk of malnutrition. About 80% of glucose deposition is mediated by skeletal muscle tissues (131,292). A young, healthy organism can rapidly convert excess glucose from dietary carbohydrates into high-molecular-weight glycogen, thereby decreasing the plasma glucose concentration to the normal level required by the central nervous system. Abnormally high plasma glucose concentrations are typically seen in elderly subjects and diabetic patients and are indicative of inadequate muscular glycogen synthase activity (131, 256, 598,650). This enzyme is regulated by physical activity and strongly enhanced by endurance exercise (40,58, 597, 598), suggesting that the sedentary life-style of the elderly may contribute to an age-related increase in systemic mean glucose level.

G.  ROS Production in Skeletal Muscle Tissue During Physical Immobilization and Intensive Physical Exercise

ROS production was reported to increase in skeletal muscle tissue after immobilization (314, 315). In view of the massive loss of skeletal muscle mass and body cell mass in immobilized subjects (110, 202,259, 262, 540,583), this finding certainly deserves more intensive investigation and underscores the importance of muscular activity.

However, intensive muscular activity was also reported to enhance ROS production (129, 382, 383,534). With the use of electron spin resonance, it was found that free radical concentrations were increased more than twofold in rat skeletal muscle and liver tissues after exhaustive exercise (129, 280). Repetitive muscular contraction was reported to increase ROS levels and thereby promote the low-frequency fatigue of muscle fibers (459). Direct evidence for increased rates of ROS production during intensive physical exercise is still scarce but is supported by other putative manifestations of oxidative stress, such as changes in the thiol/disulfide redox state in blood plasma and erythrocytes, in the context of intensive physical exercise. A decrease in intracellular GSH/GSSG ratios was found in skeletal muscle of rats after intense muscular exercise (348, 511); a similar decrease in GSH/GSSG ratios was detected in the blood of human volunteers after strenuous exercise (164, 201, 485, 513). The thiol compound N-acetylcysteine was found to ameliorate muscle fatigue in humans (460).

The rate of ROS production is generally believed to increase during intensive physical exercise as a consequence of the increased oxygen consumption of the exercising muscle. However, xanthine oxidase may also contribute to the production of superoxide in the context of intensive physical exercise. After exhaustive physical exercise, patients with chronic obstructive pulmonary disease exhibit a significant increase in the GSSG/GSH ratio in the arterial blood and a significant increase in malondialdehyde, a product of lipid peroxidation (253). These symptoms are ameliorated by treatment with allopurinol, a potent inhibitor of xanthine oxidase. A marked increase in xanthine oxidase activity was also found in the plasma of rats after exhaustive exercise (599). Treatment with allopurinol ameliorated the oxidative shift in the glutathione redox state of the blood and the appearence of the cytosolic enzymes creatine kinase and aspartate aminotransferase in the plasma of human volunteers and rats after exhaustive physical exercise (599). Under normal conditions, xanthine oxidase accounts for only a minor proportion of total ROS production (reviewed in Ref.101).

H.  Oxidative Stress as a Frequent Complication in Disease Conditions

There is a growing awareness that oxidative stress plays a role in various clinical conditions. Malignant diseases, diabetes, atherosclerosis, chronic inflammation, human immunodeficiency virus (HIV) infection, ischemia-reperfusion injury, and sleep apnea are important examples. These diseases fall into two major categories. In the first category, diabetes mellitus and cancer show commonly a pro-oxidative shift in the systemic thiol/disulfide redox state and impaired glucose clearance, suggesting that skeletal muscle mitochondria may be the major site of elevated ROS production (see sect. vi F). These conditions may be referred to as "mitochondrial oxidative stress." Without therapeutic intervention these conditions lead to massive skeletal muscle wasting, reminiscent of aging-related wasting. The second category may be referred to as "inflammatory oxidative conditions" because it is typically associated with an excessive stimulation of NAD(P)H oxidase activity by cytokines or other agents. In this case increased ROS levels or changes in intracellular glutathione levels are often associated with pathological changes indicative of a dysregulation of signal cascades and/or gene expression, exemplified by altered expression of cell adhesion molecules (120,152, 325, 405, 438,574).

I.  Malignant Diseases

ROS are potential carcinogens because they facilitate mutagenesis, tumor promotion, and progression (155, 223,269, 410, 483). The growth-promoting effects of ROS are related to redox-responsive signaling cascades, some of which have been discussed in sectionv. A placebo-controlled clinical study of patients with previous adenomatous colonic polyps, i.e., a group with an increased risk for colon cancer and increased proliferative index of colonic crypts, revealed a significant decrease in the proliferative index after treatment with N-acetylcysteine (170). Even normal cells often show increased proliferation and expression of growth-related genes if exposed to hydrogen peroxide or superoxide (81, 118, 123,419). In addition, certain types of cancer cells produce substantial amounts of ROS (36, 223,387, 553, 557, 562,653). ROS production is induced after the expression of several genes associated with a transformed phenotype including H-Rasv12 or mox1 (276,553). Ras transformation of fibroblasts and the induction of ROS production were shown to involve Rac1 and may thus be similar to the pathway by which extracellular growth factors induce ROS production (see sects. iv A3 and v A). The apparent inconsistency between the uncontrolled cell growth in ROS-producing malignant cells and the ROS-induced senescence in normal cells suggests, however, that ROS production may be necessary but not sufficient to induce malignant cell growth (see sect.vi D).

A pro-oxidative shift in the plasma thiol/disulfide redox state has been observed in patients with various types of advanced malignancies (225). This shift is reminiscent of similar changes in diabetes mellitus, old age, and intensive physical exercise. Because cancer patients commonly have decreased glucose clearance capacity (371, 414, 447,469, 568) and, in addition, abnormally high glycolytic activity and lactate production (517,518, 568, 611), it is reasonable to assume that the observed pro-oxidative shift is mediated by an increased and basically uncontrolled availability of mitochondrial energy substrate.

J.  Diabetes Mellitus

Elevated ROS levels have also been implicated in diabetes mellitus (47, 629). In this case oxidative stress is associated with a pro-oxidative shift of the glutathione redox state in the blood (137). Hyperglycemia is a hallmark of both non-insulin-dependent (type 2) and insulin-dependent diabetes mellitus (type 1). Elevated glucose levels are associated with increased production of ROS by several different mechanisms (47, 415, 420, 482,593). In cultured bovine aortic endothelial cells, hyperglycemia was shown to cause increased ROS production at the mitochondrial complex II (415). Several independent strategies that ameliorate mitochondrial ROS production were shown to prevent some of the typical secondary complications of the disease, including the activation of protein kinase C or NF-κB and the formation of advanced glycation end products (415).

In addition, superoxide is generated by the process of glucose auto-oxidation that is associated with the formation of glycated proteins in the plasma of diabetic patients (47,420, 482, 593, 628,630). The interaction of advanced glycation end products with corresponding cell surface receptors stimulates ROS production and decreases intracellular glutathione levels (645). The increase in ROS production contributes to the development of diabetic complications such as atherosclerosis and other vascular complications (25, 47). In addition, hyperglycemia enhances cell-mediated low-density lipoprotein (LDL) peroxidation in endothelial cells (381). Treatment with antioxidants ameliorates diabetic complications including the dysfunction of endothelial cells or increased platelet aggregation (90,99, 271, 298, 364,528, 626).

K.  Atherosclerosis

Atherosclerosis is a multifactorial disease characterized by hardening and thickening of the arterial wall. The vascular areas affected by this disease contain mononuclear cells, proliferating smooth muscle cells, and extracellular matrix components. Atherosclerosis is commonly viewed as a chronic inflammatory disease and is associated with certain risk factors such as hyperlipidemia, diabetes (see sect. vi J), and hypertension. Excessive ROS production has been implicated in the pathogenesis of atherosclerosis and hypertension (19, 29,98, 210, 227, 328,547, 550). Oxidative stress induces the expression of protein kinases such as focal adhesion kinase and intercellular adhesion molecules such as ICAM-1 (106). The invasion of the artery wall by monocytes and T lymphocytes is one of the earliest events in the development of atherosclerotic lesions. Monocytes, macrophages, and smooth muscle cells possess the so-called scavenger receptor for oxidized LDL. Binding of oxidized LDL leads to the activation of monocytes and macrophages and stimulates the expression of Mn-SOD which, in turn, increases the concentration of hydrogen peroxide by perturbing the steady-state levels of ROS (305, 306, 308). This process is associated with massive macrophage apoptosis and contributes thereby to the formation of the atherosclerotic lesions (309, 461). The process may be further enhanced by cytokines and other factors such as TNF, interleukin-1β, angiotensin II, and interferon-γ, which induce superoxide production by the membrane-bound NADPH oxidase in endothelial cells (136, 210, 396,424).

Lipid peroxidation and atherogenesis may be ameliorated by vitamin E. A study of atherosclerosis-susceptible APO-lipoprotein E knock-out mice revealed that induction of vitamin E deficiency by disruption of the α-tocopherol transfer protein gene (Ttpa) increased the severity of atherosclerotic lesions in the proximal aorta (570). A randomized placebo-controlled study of patients with angiographically proven coronary artery disease revealed that NO-dependent, flow-mediated dilation was significantly improved by treatment with the cysteine pro-drug and glutathione precursor l-2-oxo-4-thiazolidine carboxylate (600).

In view of the enhancement of T-cell signaling by oxidative conditions (see sect. iv F), it is not surprising that the atherosclerotic process also shows signs of an autoimmune component (618, 619). T cells isolated from atherosclerotic lesions in rabbits receiving a cholesterol-rich diet were found to express a preferential reactivity toward the mycobacterial heat shock protein hsp65, which shows >50% sequence homology with human hsp60 on the DNA and protein levels (641). A role of hsp65/60 in association with oxidative stress has also been implicated in other autoimmune diseases such as rheumatoid arthritis (303, 462,594) (see also sect. vi M).

L.  Neurodegenerative Diseases

Down's syndrome or trisomy 21 is the most frequent genetic cause of mental retardation and is commonly associated with the development of Alzheimer's disease (AD) in adult life. Cultured cortical neurons from fetal Down's syndrome cases exhibit a three- to-fourfold higher intracellular ROS level than age-matched normal brain cells. Treatment with free radical scavengers or catalase prevents the degeneration of Down's syndrome neurons in culture (84). The gene of the antioxidant enzyme Cu/Zn-SOD is localized at 21q22.1 (529) and, through increased gene dosage, is thought to play a prominent role in some of the clinical features of Down's syndrome. Cu/Zn-SOD was found to be elevated in a variety of cell types and organs including erythrocytes, platelets, fibroblasts, lymphocytes, and fetal brain (24, 73, 133, 135, 180; reviewed in Ref. 55). This increase in SOD is expected to result in increased generation of hydrogen peroxide and a displaced equilibrium in the steady-state levels of ROS.

Stably transfected cells overexpressing Cu/Zn-SOD show a higher level of lipid peroxidation than control cells (168). Cells exhibiting an increase in the ratio of Cu/Zn-SOD to glutathione peroxidase and in catalase activity also have a higher intracellular level of hydrogen peroxide and show typical features of cellular senescence (see sect. vi D), as indicated by a slower growth rate and altered morphology (134). A similar senescent phenotype is observed in cells from children with Down's syndrome and in cells exposed directly to hydrogen peroxide (56, 134). Cu/Zn-SOD transgenic mice also exhibit signs of premature aging and neuromuscular dysfunction (30-32, 648). The neuromuscular junctions of the leg muscles from these mice were found to exhibit pathological changes that are similar to those observed in the skeletal muscle tissue of aging rats and mice (95, 172) and in the tongue muscles of individuals with Down's syndrome (646, 647).

AD is a neurodegenerative disorder characterized by a progressive decline in cognitive function and extensive neuronal loss. The brains of affected patients show numerous amyloid plaques and neurofibrillary tangles. The production of ROS in the brains of AD patients and its implication in AD pathogenesis are implicated by the significant amount of lipid peroxidation detected in the brain as well as by the increased levels of 4-hydroxynonenal found in postmortem cerebrospinal fluid of AD patients (363, 398, 404,450, 487). Furthermore, ROS were found to mediate amyloid β-protein damage (51, 293,404).

Amyotrophic lateral sclerosis (ALS) is a neurodegenerative disease that affects primarily motor neurons in the spinal cord and brain stem. Approximately 10% of the cases are inherited in an autosomal dominant manner. One-fifth of these familial ALS patients carry mutations in the Cu/Zn-SOD gene, suggesting the involvement of ROS in this neurodegenerative disease (470). Several lines of transgenic mice carrying mutant SOD transgenes have been shown to develop a pathology and clinical phenotype similar to that of familial ALS patients (582). The mutation in the Cu/Zn-SOD gene causes neuronal death by apoptosis through the sequential activation of caspase-1 and caspase-3 (440).

Transmissible spongiform encephalopathies (TSEs) are characterized by the conversion of the cellular form of the prion protein (PrPC) into a conformationally modified protease-resistant isoform called PrPSc(451). A prominent example of such a prion disease is the bovine spongiform encephalopathy (BSE). The function of PrPc and particularly its role in the neurodegenerative processes, which are typically found in prion diseases, are still unknown. It has been proposed, however, that PrPC may play a role in the control of the oxidative state of the cell through a regulation of the copper transport (441) and/or through a modification of Cu/Zn-SOD activity (74). More recently, Milhavet et al. (392) demonstrated that prion-infected neuronal cells displayed a higher sensitivity to oxidative stress over noninfected cells. Infected cells also showed increased lipid peroxidation and a dramatic decrease in various redox-related enzymes such as glutathione peroxidase, glutathione reductase, and Mn-SOD (392). Collectively, these findings suggest that prion infection compromises the cellular resistance to ROS.

M.  Rheumatoid Arthritis

While the enhancement of immune reactivity by pro-oxidative conditions may be critically important for the immune system to control and defeat rapidly multiplying pathogens (see sect.iv F), such enhancement also bears the risk of inducing autoimmune processes. Rheumatoid arthritis is a systemic autoimmune disease characterized by chronic joint inflammation with infiltration of macrophages and activated T cells. Several lines of evidence suggest that production of free radicals at the site of inflammation may contribute decisively to the pathogenesis of this disease (26, 370, 379). T cells isolated from the synovial fluid of patients with rheumatoid arthritis are characterized by a decreased intracellular GSH level and the "primed" CD45RO phenotype (378). These T cells exhibit severely impaired phosphorylation of the adaptor protein linker for T-cell activation (LAT). Changes in intracellular GSH level were shown to alter the subcellular localization of LAT (214). The migration of monocytes and lymphocytes into the rheumatoid arthritis synovium is mediated by the abnormal expression of several adhesion molecules including ELAM-1, VCAM-1, ICAM-1, and ICAM-2 (119, 120, 595), an effect which may be explained by the abnormal induction of redox-sensitive signaling pathways (see sect. iv E). Oxidative conditions in synovial tissue are also associated with a higher incidence of p53 mutations (179). Although malignant tumors of the synovium are rare, it has been hypothesized that the presence of transformed cells in the synovium of rheumatoid arthritis patients may lead to progressive joint destruction without malignant degeneration (563). The heat shock protein hsp65/60, which has been implicated in the pathogenesis of atherosclerosis (see sect.vi K), is also a candidate (auto)antigen in the pathogenesis of rheumatoid arthritis (303,462, 594).

N.  HIV Infection

HIV infection is associated with progressive deterioration of the immune system, leading eventually to lethal opportunistic infections. Relatively early in the course of HIV infection, there is a decrease in various functional activities of lymphoid cells followed by a conspicuous decrease in CD4+ T-lymphocyte numbers. In the late stages of the disease, the patients often suffer from massive skeletal muscle wasting.

HIV infection is also associated with massive catabolism of cysteine into sulfate. This process can be detected even in the early asymptomatic state of the disease and accounts for a mean net loss of more than 4 g cysteine/day (70, 156). Excessive cysteine catabolism can be detected most easily as urinary sulfate excretion or as muscular sulfate excretion determined from arterial-venous differences in the lower extremities (70). The ratio of urinary sulfate to urea (i.e., sulfur/nitrogen) indicates that the excessive cysteine catabolism proceeds largely at the expense of glutathione rather than protein. This is in agreement with studies of SIV-infected rhesus macaques which exhibit a progressive decrease in the glutathione level of skeletal muscle tissue followed by skeletal muscle wasting (216). The mechanism responsible for excessive cysteine catabolism is unclear. Old age and intensive physical exercise, i.e., two conditions believed to be associated with increased muscular ROS production, show typically increased rather than decreased plasma cystine (cysteine disulfide) levels. In contrast, HIV-infected patients and SIV-infected rhesus macaques show, on the average, abnormally low plasma cystine levels (157,158, 166, 225, 261) and low intracellular glutathione levels in peripheral blood lymphocytes (165, 468).

The limiting role of glutathione and its precursor cysteine in the execution of the immune response was discussed in sectionv J. Numerous lymphocyte functions are exquisitely sensitive to a decrease in intracellular glutathione levels (reviewed in Ref. 160). That the virus-induced cyst(e)ine deficiency is indeed a causative factor for the progressive impairment of the immune system is suggested by two independent placebo-controlled double-blind studies, which demonstrated that treatment of HIV-infected patients with N-acetylcysteine leads to a significant improvement of various proliferative T-cell responses and to a reconstitution of NK cell activity to almost normal levels (71). In view of the importance of glutathione levels in several redox-regulated systems (see sect.v J), it is believed that the HIV-induced decrease in intracellular glutathione levels facilitates the induction of signaling pathways leading to lymphocyte activation but renders the cells more sensitive to oxidative stress. Ex vivo labeling studies have shown that, compared with healthy controls, HIV-infected patients have indeed significant increases in the number and faction of dividing CD4+ and CD8+ T cells. The fact that CD4+ T-cell counts decline during the course of HIV infection suggests, however, that the increase in CD4+ T cell destruction is greater than the increase in T-cell production (345). Taken together, the available evidence suggests that depletion of the systemic cysteine pool may be one of several ways by which a virus can prevent its elimination by the immune system. Because immune reconstitution is a widely accepted aim of HIV therapy, cysteine supplementation may be considered as a standard therapy for these patients.

O.  Ischemia and Reperfusion Injury

Ischemia and reperfusion can lead to tissue injury and are serious complications in organ transplantation, myocardial infarction, and stroke (194, 196, 203). Massive ROS production was identified as an important causative factor (100, 116, 153,282, 375, 376, 385,432, 433, 579). Xanthine dehydrogenase, which normally utilizes NAD+ as electron acceptor, is converted under the conditions of ischemia/reperfusion into xanthine oxidase which uses oxygen as substrate. During the ischemic period, excessive ATP consumption leads to the accumulation of the purine catabolites hypoxanthine and xanthine, which upon subsequent reperfusion and influx of oxygen are metabolized by xanthine oxidase to yield massive amounts of superoxide and hydrogen peroxide (208). More recently, a Rac1-regulated NAD(P)H oxidase distinct from the phagocytic NAD(P)H oxidase was shown to be critically involved in ROS production in a mouse model of hepatic ischemia/reperfusion injury (434).

Neutrophils are the principal effector cells of reperfusion injury, and the inhibition of neutrophil adhesion to the endothelium attenuates the process (573). Antioxidant treatment ameliorates both leukocyte adhesion and leukocyte-mediated heart injury in the postischemic period (514). Also, treatment with a synthetic SOD mimetic was shown to ameliorate tissue damage in a rat model of ischemia/reperfusion injury (484).

Experimental induction of ischemia and reperfusion in the rat heart was found to be associated with the activation of the redox-responsive trancription factors NF-κB and AP-1 and the MAPKs JNK and p38 in the presence of minimal activation of ERK (112,341, 377). This activation may account for inflammatory responses and apoptotic cell death in the affected tissue (296, 502).

P.  Obstructive Sleep Apnea

A substantial proportion of the adult population in Western countries suffers from a breathing disorder characterized by repeated episodes of apnea or hypopnea during sleep (652). This condition is associated with the development of hypertension (444). In its more severe manifestations, obstructive sleep apnea is associated with increased mortality due to cardiovascular morbidity, including arterial hypertension, coronary artery disease, and cerebrovascular disease (244,443, 519). It is one of the most important cardiovascular risk factors.

The resulting repetitive hypoxia/reoxygenation stress is reminiscent of the ischemia/reperfusion condition. An involvement of ROS in the pathogenesis of cardiovascular complications is suggested by the finding that polymorphonuclear neutrophils from the blood of patients with obstructive sleep apnea show significantly increased superoxide production after exposure to different stimuli (503). In addition, obstructive sleep apnea is associated with increased expression of several cell adhesion molecules such as ICAM-1 and VCAM-1, suggesting that the pathogenesis of the cardiovascular complications are mechanistically similar to the development of atherosclerosis (see sect. vi K) (425).

Q.  Key Messages From Section vi

The important role of ROS in the regulation of physiological responses is underscored by the apparent dysregulation of physiological responses in various disease-related oxidative stress conditions.

Excessive levels of ROS may be generated either by excessive stimulation of otherwise tightly regulated NAD(P)H oxidases or by other mechanisms that produce ROS "accidentally" in a nonregulated fashion. The latter situation includes the production of ROS by the mitochondrial ETC, the quantitatively most important source of ROS in higher organisms.

ROS production in the mitochondrial ETC is determined in part by substrate availability. In endothelial cells elevation of the extracellular glucose concentration was shown to cause a massive increase in ROS production by the mitochondrial ETC and a corresponding activation of NF-κB.

Caloric restriction is an effective experimental strategy to decrease oxidative stress in vivo and to increase the life span of experimental animals. Endurance exercise can serve to some extent as a substitute for caloric restriction in humans.

Intensive physical exercise induces various manifestations of oxidative stress.

Diabetes mellitus and cancer commonly show a pro-oxidative shift in the systemic thiol/disulfide redox state, similar to the shift seen in old age. These "mitochondrial oxidative stress" conditions are typically associated with skeletal muscle wasting. In diabetes mellitus ROS may also be generated by glucose auto-oxidation, which is associated with the formation of glycated proteins in the plasma. ROS are potential carcinogens because they facilitate mutagenesis, tumor promotion, and progression. The growth-promoting effects of ROS are related to redox-responsive signaling cascades.

Another group of diseases designated as "inflammatory oxidative conditions" includes atherosclerosis and chronic inflammation and is typically associated with excessive stimulation of NAD(P)H oxidase activity. Some of the pathological complications are indicative of a redox-mediated dysregulation of signaling cascades and/or gene expression, e.g., the overproduction of cell adhesion molecules.

ROS production by xanthine oxidase has been implicated in the pathology of ischemia and reperfusion injury. Similar mechanisms may also operate in obstructive sleep apnea.

The free radical theory of aging (234) states that age-related degenerative processes are to a large extent the consequence of damage induced by free radicals. A growing body of evidence now suggests that aging involves, in addition, progressive changes in free radical-mediated regulatory processes, resulting in altered gene expression.

There are various indirect manifestations of oxidative stress in old age, including lipid peroxidation, DNA oxidation, protein oxidation, and a shift in the redox states of thiol/disulfide redox couples such as glutathione, cysteine, and albumin. All of these manifestations suggest that the rate of ROS production per time unit increases with age. However, this conclusion needs to be tested experimentally. ROS production is difficult to measure in biological tissues.

Regardless of whether the documented age-related shift in the thiol/disulfide redox state is caused by increasing ROS levels or not, this shift itself may cause a progressive redox-mediated dysregulation of signaling cascades (see sect.iii A3).

The term replicative senescence describes the fact that somatic cells divide in cell culture only for a finite number of generations. Recent evidence suggests that the decrease in replicative capacity may be related to an increase in ROS levels and/or a pro-oxidative shift in the thiol/disulfide redox state. Mild oxidative stress was also shown to cause telomere shortening in fibroblasts.

VII. CONCLUSIONS

A.  Physiological Aspects of Redox Regulation

The radicals NO (NO·) and superoxide anion (O 2 ·) play an important role in biological regulation. Superoxide gives rise to other forms of ROS that serve as mediators in many regulatory processes. Most redox-responsive regulatory mechanisms in bacteria and mammalian cells serve to protect the cells against oxidative stress and to reestablish redox homeostasis. The oxidative induction of protective enzymes by the redox-sensitive bacterial OxyR and SoxR proteins or the inhibition of NOS by NO are prominent examples. Redox regulation of other physiological responses in higher organisms is embedded in these basic mechanisms of redox homeostasis.

The relatively large number of NAD(P)H oxidase isoforms and NOS indicates that nature has "learned" to use free radicals to her advantage in processes not directly related to protection against oxidative stress. The production of superoxide and NO, respectively, by these enzymes is strictly regulated by hormones, cytokines, or other inducing mechanisms. The resulting oxidative species, in turn, act as secondary messengers to control a variety of physiological responses. The regulation of vascular smooth muscle relaxation, the monitoring of the oxygen concentration in the regulation of respiratory ventilation and erythropoietin production, and the enhancement of signaling cascades from various membrane receptors are prominent examples.

The enhancement of signal transduction from a given receptor by stimulation of ROS production through this or other receptors may serve two physiological purposes. First, it provides a basis for cooperativity, and second, the membrane receptor may function simultaneously as a sensor for the extracellular ligand and as a sensor for the inner metabolic state of the individual cell. The cooperativity between the angiotensin II receptor and the EGF receptor is a well-studied example, but other examples will likely be found. Because hydrogen peroxide has a relatively long half-life and crosses membranes, the cooperativity principle may even extend to other cells in the vicinity. By enhancing the intracellular signaling pathways of lymphocytes, ROS from activated macrophages and neutrophils may contribute decisively to the activation of the antigen-specific immune response and may allow the immune system to respond to minute amounts of invading pathogens. Signaling pathways involving JNK, p38 MAPK, and the transcription factors AP-1 and NF-κB are particularly responsive to redox regulation.

B.  Molecular Aspects of Redox Regulation: Gain of Function, Loss of Function, or Outright Destruction

The capacity of ROS to damage proteins and to hasten their proteolytic degradation has been employed as a regulatory mechanism in several cases, e.g., in the degradation of the transcription factor subunit HIF-1α and the NF-κB inhibitor IκB. The inhibition of protein tyrosine phosphatases is well-defined on a molecular basis and provides an example of redox regulation by loss of function. In other cases, NO or ROS induce a gain of function in a signaling protein. This mechanism is involved in the regulation of vascular tone and the functional activation of the bacterial OxyR and SoxR proteins. The oxidative enhancement of membrane receptor signaling and the corresponding downstream signaling pathways are not well-characterized at the molecular level but are likely to involve the simultaneous induction of several different redox-sensitive signaling proteins. This redundancy does not preclude selective effects. The in vivo relevance of redox-sensitive signaling cascades is strongly suggested by the mere existence of the many NAD(P)H oxidase isoforms (see sect. iv A) and by the apparent dysregulation of physiological responses in various disease-related oxidative stress conditions (see sect.vi). However, the relative contributions of individual redox-sensitive signaling proteins to redox-regulated processes in vivo are presently obscure.

C.  Regulated Versus Uncontrolled Free Radical Production: Increased ROS Levels in Old Age and Disease

There is evidence that ROS production may be substantially elevated in old age and certain disease conditions. Excessive stimulation of NAD(P)H oxidase by cytokines or other mediators is implicated in various disease conditions. Other sources of superoxide such as the mitochondrial ETC and xanthine oxidase are not tightly regulated and may become increasingly relevant in old age, diabetes mellitus, and malignant diseases. With respect to the free radical theory of aging, we are seeing a shift in paradigm. Changes in redox-responsive signaling cascades and in the expression of corresponding target genes may have a similar or even greater impact on senescence as the direct radical-inflicted damage of cellular constituents.

Receptor signaling was also found to be strongly influenced by the intracellular glutathione level in all cases where appropriate experiments were performed. Many redox-sensitive signaling cascades respond equally well to ROS or to changes in the intracellular thiol/disulfide redox state. The massive oxidative shift in the human plasma thiol/disulfide redox state between the 3rd and the 10th decade of life may, therefore, alter the set point of redox-sensitive signaling pathways in various somatic cells. The pro-oxidative shift may account for age-related immunological dysfunctions and inflammatory processes as well as for the loss of the replicative capacity of fibroblasts, as illustrated schematically in Figure13.

Fig. 13.

Fig. 13.Oxidative stress and senescence: observed age-related changes and putative mechanisms. A persistent increase in ROS production either by mitochondria in skeletal muscle tissue or by chronic stimulation of NAD(P)H oxidase activity in leukocytes may cause a dysregulation of redox-sensitive signaling pathways in addition to direct oxidative damage. Progressive changes in the systemic plasma thiol/disulfide redox state may affect changes in redox-sensitive signaling pathways at any anatomic site.


D.  Chances for Therapeutic Intervention and Perspectives

The development of procedures to ameliorate undesirable ROS production may be one of the central issues in research on aging and oxidative stress-related diseases in the near future. The available evidence suggests that the age-related increase in ROS production may be due, at least in part, to the age-related increase in systemic glucose levels. Rigorous caloric restriction was found to ameliorate various manifestations of oxidative stress in experimental animals but may not be feasible for humans. Flooding the system with antioxidants or the overexpression of antioxidative enzymes may be just as detrimental as excessive exposure to free radicals. To ensure ordered redox-mediated signaling, life requires a delicately balanced intermediate level of free radicals and radical-derived ROS. There are some encouraging recent reports about the use of SOD/catalase mimetics in certain experimental systems. Dietary antioxidants are widely used to ameliorate excessive oxidative stress, but scientific proof of their efficacy is scarce. There is, nevertheless, a strong possibility that the process of senescence and disease-related wasting results, at least to some extent, from a progressive shift in biochemical conditions that may not be irreversible in principle.

I am grateful to Dr. William E. Hull and Dr. Lienhard Schmitz for critically reading this manuscript, to I. Fryson for her invaluable assistance in the preparation of this manuscript, and to Bettina and Sabrina for their support. I have done my best to include all relevant publications, but some omissions were inevitable. Therefore, I apologize to those colleagues whose work has not been discussed here.

REFERENCES

  • 1 Abate C, Patel L, Rauscher FJ III, Curran T Redox regulation of Fos and Jun DNA-binding activity in vitro. Science 249 199011571161
    Crossref | PubMed | ISI | Google Scholar
  • 2 Abe J, Kusuhara M, Ulevitch RJ, Berk BC, Lee JD Big mitogen-activated protein kinase 1 (BMK1) is a redox-sensitive kinase. J Biol Chem 271 19961658616590
    Crossref | PubMed | ISI | Google Scholar
  • 3 Abe JI, Berk BC Fyn and JAK2 mediate Ras activation by reactive oxygen species. J Biol Chem 274 1999210032110
    Crossref | PubMed | ISI | Google Scholar
  • 4 Abe JI, Takahashi M, Ishida M, Lee JD, Berk BC c-Src is required for oxidative stress-mediated activation of big mitogen-activated protein kinase 1. J Biol Chem 272 19972038920394
    Crossref | PubMed | ISI | Google Scholar
  • 5 Abo A, Pick E, Hall A, Totty N, Teahan CG, Segal AW Activation of the NADPH-oxidase involves the small GTP-binding protein p21rac1. Nature 353 1991668670
    Crossref | PubMed | ISI | Google Scholar
  • 6 Abu-Soud HM, Wang J, Rousseau DL, Fukuto JM, Ignarro LJ, Stuehr DJ Neuronal NO synthase self-inactivates by forming a ferrous-nitrosyl complex during aerobic catalysis. J Biol Chem 270 19952299723006
    Crossref | PubMed | ISI | Google Scholar
  • 7 Acker H Mechanisms and meaning of cellular oxygen sensing in the organism. Respir Physiol 95 1994110
    Crossref | PubMed | Google Scholar
  • 8 Acker H, Dufau E, Huber J, Sylvester D Indications to an NADPH oxidase as a possible pO2 sensor in the rat carotid body. FEBS Lett 256 19897578
    Crossref | PubMed | ISI | Google Scholar
  • 9 Acker H, Xue D Mechanisms of O2 sensing in the carotid body in comparison with O2-sensing cells. NIPS 10 1995211216
    Google Scholar
  • 10 Adamson DC, Wildemann B, Sasaki M, Glass JD, McArthur JC, Christov VI, Dawson TM, Dawson VL Immunologic NO synthase: elevation in severe AIDS dementia and induction by HIV-1 gp41. Science 274 199619171921
    Crossref | PubMed | ISI | Google Scholar
  • 11 Adler V, Pincus MR, Minamoto T, Fuchs SY, Bluth MJ, Brandt-Rauf PW, Friedman FK, Robinson RC, Chen JM, Wang XW, Harris CC, Ronai Z Conformation-dependent phosphorylation of p53. Proc Natl Acad Sci USA 94 199716861691
    Crossref | PubMed | ISI | Google Scholar
  • 12 Adler V, Yin Z, Fuchs SY, Benezra M, Rosario L, Tew KD, Pincus MR, Sardana M, Henderson CJ, Wolf CR, Davis RJ, Ronai Z Regulation of JNK signaling by GSTp. EMBO J 18 199913211334
    Crossref | PubMed | ISI | Google Scholar
  • 13 Adler V, Yin Z, Tew KD, Ronai Z Role of redox potential and reactive oxygen species in stress signaling. Oncogene 18 199961046111
    Crossref | PubMed | ISI | Google Scholar
  • 14 Aikawa R, Komuro I, Yamazaki T, Zou Y, Kudoh S, Tanaka M, Shiojima I, Hiroi Y, Yazaki Y Oxidative stress activates extracellular signal-regulated kinases through Src and Ras in cultured cardiac myocytes of neonatal rats. J Clin Invest 100 199718131821
    Crossref | PubMed | ISI | Google Scholar
  • 15 Akhand AA, Pu M, Senga T, Kato M, Suzuki H, Miyata T, Hamaguchi M, Nakashima I Nitric oxide controls Src kinase activity through a sulfhydryl group modification-mediated Tyr-527-independent and Tyr-416-linked mechanism. J Biol Chem 274 19992582125826
    Crossref | PubMed | ISI | Google Scholar
  • 16 Alam J, Camhi S, Choi AM Identification of a second region upstream of the mouse heme oxygenase-1 gene that functions as a basal level and inducer-dependent transcription enhancer. J Biol Chem 270 19951197711984
    Crossref | PubMed | ISI | Google Scholar
  • 17 Albelda SM, Smith CW, Ward PA Adhesion molecules and inflammatory injury. FASEB J 8 1994504512
    Crossref | PubMed | ISI | Google Scholar
  • 18 Albina JE, Reichner JS Role of nitric oxide in mediation of macrophage cytotoxicity and apoptosis. Cancer Metastasis Rev 17 19983953
    Crossref | PubMed | ISI | Google Scholar
  • 19 Alexander RW Theodore Cooper Memorial Lecture. Hypertension and the pathogenesis of atherosclerosis: oxidative stress and the mediation of arterial inflammatory response: a new perspective. Hypertension 25 1995155161
    Crossref | PubMed | ISI | Google Scholar
  • 20 Allen RG, Tressini M Oxidative stress and gene regulation. Free Radical Biol Med 28 2000463499
    Crossref | PubMed | ISI | Google Scholar
  • 21 Alvarez M, Pennell RI, Meijer PJ, Ishikawa A, Dixon RA, Lamb C Reactive oxygen intermediates mediate a systemic signal network in the establishment of plant immunity. Cell 92 1998773784
    Crossref | PubMed | ISI | Google Scholar
  • 22 Ames BN, Shigenaga MK, Hagen TM Oxidants, antioxidants, and the degenerative diseases of aging. Proc Natl Acad Sci USA 90 199379157922
    Crossref | PubMed | ISI | Google Scholar
  • 23 Angel P, Karin M The role of jun, fos and the AP-1 complex in cell proliferation and transformation. Biochim Biophys Acta 1072 1991129157
    PubMed | ISI | Google Scholar
  • 24 Anneren G, Epstein CJ Lipid peroxidation and superoxide dismutase-1 and glutathione peroxidase activities in trisomy 16 fetal mice and human trisomy 21 fibroblasts. Pediatr Res 21 19878892
    Crossref | PubMed | ISI | Google Scholar
  • 25 Aragno M, Parola S, Tamagno E, Brignardello E, Manti R, Danni O, Boccuzzi G Oxidative derangement in rat synaptosomes induced by hyperglycaemia: restorative effect of dehydroepiandrosterone treatment. Biochem Pharmacol 60 2000389395
    Crossref | PubMed | ISI | Google Scholar
  • 26 Araujo V, Arnal C, Boronat M, Ruiz E, Dominguez C Oxidant-antioxidant imbalance in blood of children with juvenile rheumatoid arthritis. Biofactors 8 1998155159
    Crossref | PubMed | ISI | Google Scholar
  • 27 Archer SL, Huang JMC, Hampl V, Nelson DP, Shultz PJ, Weir EK Nitric oxide and cGMP cause vasorelaxation by activation of a charybdotoxin-sensitive K-channel by cGMP-dependent protein kinase. Proc Natl Acad Sci USA 91 199475837587
    Crossref | PubMed | ISI | Google Scholar
  • 28 slund 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 199961616165
    Crossref | PubMed | ISI | Google Scholar
  • 29 Auch-Schwelk W, Bossaller C, Claus M, Graf K, Grafe M, Fleck E Local potentiation of bradykinin-induced vasodilation by converting-enzyme inhibition in isolated coronary arteries. J Cardiovasc Pharmacol 9 Suppl 1992S62S67
    Google Scholar
  • 30 Auralean KB, Schickler M, Sapoznikov D, Yasom 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 1988823829
    Crossref | PubMed | ISI | Google Scholar
  • 31 Avraham K, Schickler M, Sapoznikow R, Yarom R, Groner Y Down's syndrome: abnormal neuromuscular junction in tongue of transgenic mice with levels of human CuZn superoxide dismutase. Cell 54 1988823829
    Crossref | PubMed | ISI | Google Scholar
  • 32 Avraham K, Sugaman 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 1991208215
    Crossref | PubMed | Google Scholar
  • 33 Axline SG Functional biochemistry of the macrophage. Semin Hematol 7 1970142160
    PubMed | ISI | Google Scholar
  • 34 Azzi A, Boscoboinik D, Marilley D, Özer NK, Stäuble B Vitamin E: a sensor and an information transducer of the cell oxidation state. Am J Clin Nutr 62 Suppl 19951337S1346S
    Crossref | ISI | Google Scholar
  • 35 Baas AS, Berk BC Differential activation of mitogen-activated protein kinases by H2O2 and O 2 in vascular smooth muscle cells. Circ Res 77 19952936
    Crossref | PubMed | ISI | Google Scholar
  • 36 Bae YS, Kang SW, Seo MS, Baines IC, Tekle E, Chock PB, Rhee SG Epidermal growth factor (EGF)-induced generation of hydrogen peroxide. J Biol Chem 272 1997217221
    Crossref | PubMed | ISI | Google Scholar
  • 37 Bae YS, Sung JY, Kim OS, Kim YJ, Hur KC, Kazlauskas A, Rhee SG Platelet-derived growth factor-induced H2O2 production requires the activation of phosphatidylinositol 3-kinase. J Biol Chem 275 20001052710531
    Crossref | PubMed | ISI | Google Scholar
  • 38 Baeuerle PA, Baltimore D NF-κB: ten years after. Cell 87 19961320
    Crossref | PubMed | ISI | Google Scholar
  • 39 Baeuerle PA, Henkel T Function and activation of NF-κB in the immune system. Annu Rev Immunol 12 1994141179
    Crossref | PubMed | ISI | Google Scholar
  • 40 Bak JF, Pedersen O Exercise-enhanced activation of glycogen synthase in human skeletal muscle. Am J Physiol Endocrinol Metab 258 1990E957E963
    Link | ISI | Google Scholar
  • 41 Banki K, Hutter E, Gonchoroff NJ, Perl A Elevation of mitochondrial transmembrane potential and reactive oxygen intermediate levels are early events and occur independently from activation of caspases in Fas signaling. J Immunol 162 199914661479
    PubMed | ISI | Google Scholar
  • 42 Barford D, Jia Z, Tonks NK Protein tyrosine phosphatases take off. Nat Struct Biol 2 199510431053
    Crossref | PubMed | Google Scholar
  • 43 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 19993454334546
    Crossref | PubMed | ISI | Google Scholar
  • 44 Barrett WC, DeGnore JP, Konig S, Fales HM, Keng YF, Zhang ZY, Yim MB, Chock PB Regulation of PTP1B via glutathionylation of the active site cysteine 215. Biochemistry 38 199966996705
    Crossref | PubMed | ISI | Google Scholar
  • 45 Bauer CE, Elsen S, Bird TH Mechanisms for redox control of gene expression. Annu Rev Microbiol 53 1999495523
    Crossref | PubMed | ISI | Google Scholar
  • 46 Bauss F, Dröge W, Männel DN Tumor necrosis factor mediates endotoxic effects in mice. Infect Immun 55 198716221625
    Crossref | PubMed | ISI | Google Scholar
  • 47 Baynes JW Role of oxidative stress in development of complications in diabetes. Diabetes 40 1991405412
    Crossref | PubMed | ISI | Google Scholar
  • 48 Bayreuther K, Rodemann HP, Hommel R, Dittmann K, Albiez M, Rancz PI Human skin fibroblasts in vitro differentiate along a terminal cell lineage. Proc Natl Acad Sci USA 85 198851126116
    Crossref | PubMed | ISI | Google Scholar
  • 49 Beckman KB, Ames BN The free radical theory of aging matures. Physiol Rev 78 1998547581
    Link | ISI | Google Scholar
  • 50 Beg AA, Baltimore D An essential role for NF-κB in preventing TNF-α-induced cell death. Science 274 1996782784
    Crossref | PubMed | ISI | Google Scholar
  • 51 Behl C, Davis JB, Lesley R, Schubert D Hydrogen peroxide mediates amyloid β protein toxicity. Cell 77 1994817827
    Crossref | PubMed | ISI | Google Scholar
  • 52 Beiqing L, Chen M, Whisler RL Sublethal levels of oxidative stress stimulate transcriptional activation of c-jun and suppress IL-2 promoter activation in Jurkat T cells. J Immunol 157 1996160169
    PubMed | ISI | Google Scholar
  • 53 Berg AA, Finco TS, Nantermet PV, Baldwin AS Tumor necrosis factor and interleukin-1 lead to phosphorylation and loss of IκB: a mechanism for NF-κB activation. Mol Cell Biol 13 199333013310
    Crossref | PubMed | ISI | Google Scholar
  • 54 Berlett BS, Levine RL, Stadtman ER Comparison of the effects of ozone on the modification of amino acid residues in glutamine synthetase and bovine serum albumin. J Biol Chem 271 199641774182
    Crossref | PubMed | ISI | Google Scholar
  • 55 Bladier C, de Haan JB, Kola I Antioxidant genes and reactive oxygen species in Down's syndrome. Antioxidant and Redox Regulation of Genes , Sen CK, Sies H, Baeuerly PA 2000AcademicOrlando, FL, p.425–449.
    Google Scholar
  • 56 Bladier C, Wolvetang EJ, Hutchinson P, de Haan JB, Kola I Response of a primary human fibroblast cell line to H2O2: senescence-like growth arrest or apoptosis? Cell Growth Differ 8 1997589598
    PubMed | Google Scholar
  • 57 Blumberg P Complexities of the protein kinase C pathway. Mol Carcinog 4 1991339344
    Crossref | PubMed | ISI | Google Scholar
  • 58 Bogardus C, Thuillez P, Ravussin E, Vasquez B, Narimiga M, Azhar S Effect of muscle glycogen depletion on in vivo insulin action in man. J Clin Invest 72 198316051610
    Crossref | PubMed | ISI | Google Scholar
  • 59 Bogdan C, Röllinghoff M, Diefenbach A Reactive oxygen and reactive nitrogen intermediates in innate and specific immunity. Curr Opin Immunol 12 20006476
    Crossref | PubMed | ISI | Google Scholar
  • 60 Bonizzi G, Piette J, Merville MP, Bours V Cell type-specific role for reactive oxygen species in nuclear factor κB activation by interleukin-1. Biochem Pharmacol 59 2000711
    Crossref | PubMed | ISI | Google Scholar
  • 61 Borrow J, Goddard AD, Sheer D, Solomon E Molecular analysis of acute promyelocytic leukemia breakpoint cluster region on chromosome 17. Science 249 199015771580
    Crossref | PubMed | ISI | Google Scholar
  • 62 Boscoboinik D, Szewczyk A, Hensey C, Azzi A Inhibition of cell proliferation by α-tocopherol. J Biol Chem 266 199161886194
    PubMed | ISI | Google Scholar
  • 63 Boveris A, Cadenas A, Stoppani AO Role of ubiquinone in the mitochondrial generation of hydrogen peroxide. Biochem J 156 1976435444
    Crossref | PubMed | ISI | Google Scholar
  • 64 Boveris A, Chance B The mitochondrial generation of hydrogen peroxide. General properties and effect of hyperbaric oxygen. Biochem J 134 1973707716
    Crossref | PubMed | ISI | Google Scholar
  • 65 Bowie A, O'Neill LAJ Oxidative stress and nuclear factor-κB activation. Biochem Pharmacol 59 20001323
    Crossref | PubMed | ISI | Google Scholar
  • 66 Bowie AG, Moynagh PN, O'Neill LAJ Lipid peroxidation is involved in the activation of NF-κB by tumor necrosis factor but not interleukin-1 in the human endothelial cell line ECV304. J Biol Chem 272 19972594125950
    Crossref | PubMed | ISI | Google Scholar
  • 67 Bradley TM, Hidalgo E, Leautaud V, Ding H, Demple B Cysteine-to-alanine replacements in the Escherichia coli SoxR protein and the role of the [2Fe-2S] centers in transcriptional activation. Nucleic Acids Res 25 199714691475
    Crossref | PubMed | ISI | Google Scholar
  • 68 Brand KA, Hermfisse U Aerobic glycolysis by proliferating cells: a protective strategy against reactive oxygen species. FASEB J 11 1997388395
    Crossref | PubMed | ISI | Google Scholar
  • 69 Bredt DS, Hwang PM, Glatt CE, Lowenstein C, Reed RR, Synder SH 450 Reductase. Nature 351 1991714718
    Crossref | PubMed | ISI | Google Scholar
  • 70 Breitkreutz R, Holm S, Pittak N, Beichert M, Babylon A, Yodoi J, Dröge W Massive loss of sulfur in HIV infection. AIDS Res Hum Retroviruses 16 2000203209
    Crossref | PubMed | ISI | Google Scholar
  • 71 Breitkreutz R, Pittak N, Nebe CT, Schuster D, Brust J, Beichert M, Hack V, Daniel V, Edler L, Dröge W Improvement of immune functions in HIV infection by sulfur supplementation: two randomized trials. J Mol Med 78 20005562
    Crossref | PubMed | ISI | Google Scholar
  • 72 Brody JA Prospects for an ageing population. Nature 315 1985463466
    Crossref | PubMed | ISI | Google Scholar
  • 73 Brooksbank BWL, Balazs R Superoxide dismutase, glutathione peroxidase and lipid peroxidation in Down's syndrome fetal brain. Dev Brain Res 16 19843744
    Crossref | Google Scholar
  • 74 Brown DR, Besinger A Prion protein expression and superoxide dismutase activity. Biochem J 334 1998423429
    Crossref | PubMed | ISI | Google Scholar
  • 75 Brumell JH, Burkhardt AL, Bolen JB, Grinstein S Endogenous reactive oxygen intermediates activate tyrosine kinases in human neutrophils. J Biol Chem 271 199614551461
    Crossref | PubMed | ISI | Google Scholar
  • 76 Brune B, Gotz C, Messmer UK, Sandau K, Hirvonen MR, Lapetina EG Superoxide formation and macrophage resistance to nitric oxide-mediated apoptosis. J Biol Chem 272 199772537258
    Crossref | PubMed | ISI | Google Scholar
  • 77 Buck J, Derguini F, Levi E, Nakanishi K, Hämmerling U Intracellular signaling by 14-hydroxy-retro-retinol. Science 254 199116541656
    Crossref | PubMed | ISI | Google Scholar
  • 78 Buck J, Ritter G, Dannecker L, Katta V, Cohen S, Chait B, Hämmerling U Retinol is essential for growth of activated human B cells. J Exp Med 171 199016301634
    Crossref | ISI | Google Scholar
  • 79 Buga GM, Griscavage JN, Rogers NE, Ignarro LJ Negative feedback regulation of endothelial cell function by nitric oxide. Circ Res 73 1993808812
    Crossref | PubMed | ISI | Google Scholar
  • 80 Bunn H, Poyton RO Oxygen sensing and molecular adaptation to hypoxia. Physiol Rev 76 1996839885
    Link | ISI | Google Scholar
  • 81 Burdon R Superoxide and hydrogen peroxide in relation to mammalian cell proliferation. Free Radical Biol Med 18 1995775794
    Crossref | PubMed | ISI | Google Scholar
  • 82 Burke TM, Wolin MS Hydrogen peroxide elicits pulmonary arterial relaxation and guanylate cyclase activation. Am J Physiol Heart Circ Physiol 252 1987H721H732
    Link | ISI | Google Scholar
  • 83 Burns DJ, Bell RM Protein kinase C contains two phorbol ester binding domains. J Biol Chem 266 19911833018338
    PubMed | ISI | Google Scholar
  • 84 Buscigllo J, Yankner BA Apoptosis and increased generation of reactive oxygen species in Down's syndrome neurons in vitro. Nature 378 1995776779
    Crossref | PubMed | ISI | Google Scholar
  • 85 Butler AR, Glidewell C, Li MS Nitrosyl complexes of iron-sulfur cluster. Adv Inorg Chem 32 1988335392
    Crossref | ISI | Google Scholar
  • 86 Cadenas E, Boveris A, Ragan I, Stoppani AOM Production of superoxide radicals and hydrogen peroxide by NADH-ubiquinone reductase and ubiquinol-cytochrome c reductase from beef heart mitochondria. Arch Biochem Biophys 180 1977248257
    Crossref | PubMed | ISI | Google Scholar
  • 87 Cadenas E, Poderoso JJ, Antunes F, Boveris A Analysis of the pathways of nitric oxide utilization in mitochondria. Free Radical Res 33 2000747756
    Crossref | PubMed | ISI | Google Scholar
  • 88 Cahill GF Starvation in man. N Engl J Med 282 1970668675
    Crossref | PubMed | Google Scholar
  • 89 Callsen D, Pfeilschifter J, Brüne B Rapid and delayed p42/p44 MAPK activation by nitric oxide: the role of cGMP and tyrosine phosphatase inhibition. J Immunol 161 199848524858
    PubMed | ISI | Google Scholar
  • 90 Cameron NE, Cotter MA, Maxfield EK Anti-oxidant treatment prevents the development of peripheral nerve dysfunction in streptozotocin-diabetic rats. Diabetologia 36 1993299304
    Crossref | PubMed | ISI | Google Scholar
  • 91 Camhi SL, Alam J, Otterbein L, Sylvester SL, Choi AM Induction of heme oxygenase-1 gene expression by lipopolysaccharide is mediated by AP-1 activation. Am J Respir Cell Mol Biol 13 1995387398
    Crossref | PubMed | ISI | Google Scholar
  • 92 Camhi SL, Alam J, Wiegand GW, Chin BY, Choi AM Transcriptional activation of the HO-1 gene by lipopolysaccharide is mediated by 5′ distal enhancers: role of reactive oxygen intermediates and AP-1. Am J Respir Cell Mol Biol 18 1998226234
    Crossref | PubMed | ISI | Google Scholar
  • 93 Campion EW The oldest old. N Engl J Med 330 199418191820
    Crossref | PubMed | ISI | Google Scholar
  • 94 Canbolat O, Fandrey J, Jelkmann W Effects of modulators of the production and degradation of hydrogen peroxide on erythropoietin synthesis. Respir Physiol 114 1998175183
    Crossref | PubMed | Google Scholar
  • 95 Cardasis CA Ultrastructural evidence of continued reorganization at the aging (11–26 mo) rat soleus neuromuscular junction. Anat Rec 207 1983399415
    Crossref | PubMed | Google Scholar
  • 96 Castedo M, Hirsch T, Susin SA, Zamzami N, Marchetti P, Macho A, Kroemer G Sequential acquisition of mitochondrial and plasma membrane alterations during early lymphocyte apoptosis. J Immunol 157 1996512521
    PubMed | ISI | Google Scholar
  • 97 Castro L, Rodriguez M, Radi R Aconitase is readily inactivated by peroxynitrite, but not by its precursor, nitric oxide. J Biol Chem 269 19942940929415
    PubMed | ISI | Google Scholar
  • 98 Cathcart MK, McNally AK, Morel DW, Chisolm GM Superoxide anion participation in human monocyte-mediated oxidation of low-density lipoprotein and conversion of low-density lipoprotein to a cytotoxin. J Immunol 142 198919631969
    PubMed | ISI | Google Scholar
  • 99 Ceriello A, Guigliano A, Dello Russo P, Lefebvre PJ Metabolic control may influence the increased superoxide generation in diabetic serum. Diabetes Med 8 1991540542
    Crossref | PubMed | ISI | Google Scholar
  • 100 Chan PH Role of oxidants in ischemic brain damage. Stroke 27 199611241129
    Crossref | PubMed | ISI | Google Scholar
  • 101 Chance B, Sies H, Boveris A Hydroperoxide metabolism in mammalian organs. Physiol Rev 59 1979527605
    Link | ISI | Google Scholar
  • 102 Chance B, Williams GR The respiratory chain and oxidative phosphorylation. Adv Enzymol 17 195665134
    PubMed | Google Scholar
  • 103 Chandel NS, Maltepe E, Goldwasser E, Mathieu CE, Simon MC, Schumacker PT Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc Natl Acad Sci USA 95 19981171511720
    Crossref | PubMed | ISI | Google Scholar
  • 104 Chen Q, Ames BN Senescence-like growth arrest induced by hydrogen peroxide in human diploid fibroblast F65 cells. Proc Natl Acad Sci USA 91 199441304134
    Crossref | PubMed | ISI | Google Scholar
  • 105 Chen Z, Silva H, Klessing DF Active oxygen species in the induction of plant systemic acquired resistance by salicylic acid. Science 262 199318831886
    Crossref | PubMed | ISI | Google Scholar
  • 106 Chien S, Li S, Shyy YJ Effects of mechanical forces on signal transduction and gene expression in endothelial cells. Hypertension 31 1998162169
    Crossref | PubMed | ISI | Google Scholar
  • 107 Choi D, Whittier PS, Oshima J, Funk WD Telomerase expression prevents replicative senescence but does not fully reset mRNA expression patterns in Werner syndrome cell strains. FASEB J 15 200110141020
    Crossref | PubMed | ISI | Google Scholar
  • 108 Christman MF, Morgan RW, Jacobson FS, Ames BN Positive control of a regulon for defenses against oxidative stress and some heat-shock proteins in Salmonella typhimurium. Cell 41 1985753762
    Crossref | PubMed | ISI | Google Scholar
  • 109 Christman MF, Storz G, Ames BN OxyR, a positive regulator of hydrogen peroxide-inducible genes in Escherichia coli and Salmonella typhimurium, is homologous to a family of bacterial regulatory proteins. Proc Natl Acad Sci USA 86 198934843488
    Crossref | PubMed | ISI | Google Scholar
  • 110 Cintrón-Trvino NM, Leach CS, Rambaut PC Potential biochemical basis of muscle atrophy during prolonged weightlessness. Biochemistry of Exercise. Proceedings of the International Symposium on Exercise , Knuttgen HG 13 1983351355Human KineticsBoston, MA (Int Ser Sport Sci)
    Google Scholar
  • 111 Clapham DE Calcium signaling. Cell 80 1995259268
    Crossref | PubMed | ISI | Google Scholar
  • 112 Clerk A, Fuller SJ, Michael A, Sugden PH Stimulation of "stress-regulated" mitogen-activated protein kinases (stress-activated protein kinases/c-Jun N-terminal kinases and p38-mitogen-activated protein kinases) in perfused rat hearts by oxidative and other stresses. J Biol Chem 273 199872287234
    Crossref | PubMed | ISI | Google Scholar
  • 113 Cohen GM Caspases: the executioners of apoptosis. Biochem J 326 1997116
    Crossref | PubMed | ISI | Google Scholar
  • 114 Commoner B, Townsend J, Pake GE Free radicals in biological materials. Nature 174 1954689691
    Crossref | PubMed | ISI | Google Scholar
  • 115 Cooper JT, Stroka DM, Brostjan C, Palmetshofer A, Bach FH, Ferran C A20 blocks endothelial cell activation through a NFκB-dependent mechanism. J Biol Chem 271 19961806818073
    Crossref | PubMed | ISI | Google Scholar
  • 116 Cordis GA, Maulik G, Bagchi D, Riedel W, Das DK Detection of oxidative DNA damage to ischemic reperfused rat hearts by 8-hydroxydeoxyguanosine formation. J Mol Cell Cardiol 30 199819391944
    Crossref | PubMed | ISI | Google Scholar
  • 117 Crawford DR, Cerutti PA Expression of oxidant stress-related genes in tumor promotion of mouse epidermal cells JB6. Proceedings of the Second Conference on Anticarcinogens and Radioprotectors , Nygaard OF 1987183190PlenumNew York
    Google Scholar
  • 118 Crawford DR, Zbinden I, Amstad P, Cerutti P Oxidant stress induces the proto-oncogenes c-fos and c-myc in mouse epidermal cells. Oncogene 3 19882732
    ISI | Google Scholar
  • 119 Cunnane G, Hummel KM, Muller-Ladner U, Gay RE, Gay S Mechanism of joint destruction in rheumatoid arthritis. Arch Immunol Ther Exp 46 199817
    PubMed | Google Scholar
  • 120 Cutolo M, Sulli A, Barone A, Seriolo B, Accardo S Macrophages, synovial tissue and rheumatoid arthritis. Clin Exp Rheumatol 11 1993331339
    PubMed | ISI | Google Scholar
  • 121 Czech MP, Lawrence JC, Lynn WS Evidence for the involvement of sulfhydryl oxidation in the regulation of fat cell hexose transport by insulin. Proc Natl Acad Sci USA 71 197441734177
    Crossref | PubMed | ISI | Google Scholar
  • 122 Dalton TP, Shertzer HG, Puga A Regulation of gene expression by reactive oxygen. Annu Rev Pharmacol Toxicol 39 199967101
    Crossref | PubMed | ISI | Google Scholar
  • 123 Datta R, Hallahan DE, Kharbanda SM, Rubin E, Sherman ML, Huberman E, Weichselbaum RR, Kufe DW Involvement of reactive oxygen intermediates in the induction of c-jun gene transcription by ionizing radiation. Biochemistry 31 199283008306
    Crossref | PubMed | ISI | Google Scholar
  • 124 Davies KJA Protein damage and degradation by oxygen radicals. I. General aspects. J Biol Chem 262 198798959901
    PubMed | ISI | Google Scholar
  • 125 Davies KJA The broad spectrum of responses to oxidants in proliferating cells: a new paradigm for oxidative stress. Life 48 19994147
    PubMed | Google Scholar
  • 126 Davies KJA, Delsignore ME, Lin SW Protein damage and degradation by oxygen radicals. II. Modification of amino acids. J Biol Chem 262 198799029907
    PubMed | ISI | Google Scholar
  • 127 Davies KJA, Goldberg AL Oxygen radicals stimulate intracellular proteolysis and lipid peroxidation by independent mechanisms in erythrocytes. J Biol Chem 262 198782208226
    PubMed | ISI | Google Scholar
  • 128 Davies KJA, Lin SW, Pacifici RE Protein damage and degradation by oxygen radicals. IV. Degradation of denatured protein. J Biol Chem 262 198799149920
    PubMed | ISI | Google Scholar
  • 129 Davies KJA, Quintanilha AT, Brooks GA, Packer L Free radicals and tissue damage produced by exercise. Biochem Biophys Res Commun 107 198211981205
    Crossref | PubMed | ISI | Google Scholar
  • 130 Deby C, Goutier R New perspectives on the biochemistry of superoxide anion and the efficiency of superoxide dismutases. Biochem Pharmacol 39 1990399405
    Crossref | PubMed | ISI | Google Scholar
  • 131 DeFronzo RA Pathogenesis of type 2 diabetes: metabolic and molecular implications for identifying diabetes genes. Diabetes Rev 5 1997177269
    Google Scholar
  • 132 De Groot CJ, Ruuls SR, Theeuwes JW, Dijkstra CD, van der Valk P Immunocytochemical characterization of the expression of inducible and constitutive isoforms of nitric oxide synthase in demyelinating multiple sclerosis lesions. J Neuropathol Exp Neurol 56 19971020
    Crossref | PubMed | ISI | Google Scholar
  • 133 De Haan JB, Christiano F, Iannello RC, Kola I Cu/Zn-superoxide dismutase and glutathione peroxidase during aging. Biochem Mol Biol Int 35 199512811297
    PubMed | Google Scholar
  • 134 De Haan JB, Christiano F, Iannello RC, 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 1996283292
    Crossref | PubMed | ISI | Google Scholar
  • 135 De Haan JB, Wolvetang EJ, Christiano F, Iannello RC, Bladier C, Kelner MJ, Kola I Reactive oxygen species and their contribution to pathology in Down's syndrome. Adv Pharmacol 38 1997379402
    Crossref | PubMed | Google Scholar
  • 136 De Keulenaer GW, Alexander RW, Ushio-Fukai M, Ishizaka N, Griendling KK Tumor necrosis factor-α activates a p22phox-based NADH oxidase in vascular smooth muscle. Biochem J 329 1998653657
    Crossref | PubMed | ISI | Google Scholar
  • 137 De Mattia G, Bravi MC, Laurenti O, Cassone-Faldetta M, Armeinto A, Ferri C, Balsano F Influence of reduced glutathione infusion on glucose metabolism in patients with non-insulin-dependent diabetes mellitus. Metabolism 47 1998993997
    Crossref | PubMed | ISI | Google Scholar
  • 138 Demple B, Halbrook J Inducible repair of oxidative DNA damage in Escherichia coli. Nature 304 1983466468
    Crossref | PubMed | ISI | Google Scholar
  • 139 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 199856335642
    Crossref | PubMed | ISI | Google Scholar
  • 140 Deora AA, Lander HM Regulation of signal transduction and gene expression by reactive nitrogen species. Antioxidant and Redox Regulation of Genes , Sen CK, Sies H, Baeuerle PA 2000147178AcademicOrlando, FL
    Google Scholar
  • 141 Derguini F, Nakanishi K, Hämmerling U, Chua R, Eppinger T, Levy E, Buck J 13,14-Dihydroxy-retinol, a new bioactive retinol metabolite. J Biol Chem 270 19951887518880
    Crossref | PubMed | ISI | Google Scholar
  • 142 Desai VG, Weindruch R, Hart RW, Feuers RJ Influence of age and dietary restriction on gastrocnemius electron transport system activities in mice. Arch Biochem Biophys 333 1996145151
    Crossref | PubMed | ISI | Google Scholar
  • 143 De The H, Chomienne C, Lanotte M, Degos L, Dejean A The t(15;17) translocation of acute promyelocytic leukaemia fuses the retinoic acid receptor α gene to a novel transcribed locus. Nature 347 1990558561
    Crossref | PubMed | ISI | Google Scholar
  • 144 Devary Y, Gottlieb RA, Laus LF, Karin M Rapid and preferential activation of the c-jun gene during the mammalian UV response. Mol Cell Biol 11 199128042811
    Crossref | PubMed | ISI | Google Scholar
  • 145 De Vos K, Goossens V, Boone E, Vercammen D, Vancompernolle K, Vandenabeele P, Haegeman G, Fiers W, Grooten J The 55-kDa tumor necrosis factor receptor induces clustering of mitochondria through its membrane-proximal region. J Biol Chem 273 199896739680
    Crossref | PubMed | ISI | Google Scholar
  • 146 Dickens M, Tavaré JM Analysis of the order of autophosphorylation of human insulin receptor tyrosines 1158, 1162 and 1163. Biochem Biophys Res Commun 186 1992244250
    Crossref | PubMed | ISI | Google Scholar
  • 147 Dimri GP, Lee X, Basile G, Acosta M, Scott G, Roskelley C, Medrano EE, Linskens M, Rubelj I, Pereira-Smith O, Peacocke M, Campisi J A biomarker that identifies senscent human cells in culture and in aging skin in vivo. Proc Natl Acad Sci USA 92 199593639367
    Crossref | PubMed | ISI | Google Scholar
  • 148 Dinauer MC, Deck MB, Unanue ER Mice lacking reduced nicotinamide adenine dinucleotide phosphate oxidase activity show increased susceptibility to early infection with Listeria monocytogenes. J Immunol 158 199755815583
    PubMed | ISI | Google Scholar
  • 149 Ding H, Demple B In vivo kinetics of a redox-regulated transcriptional switch. Proc Natl Acad Sci USA 94 199784458449
    Crossref | PubMed | ISI | Google Scholar
  • 150 Doan TN, Gentry DL, Taylor AA, Elliott SJ Hydrogen peroxide activates agonist-sensitive Ca2+-flux pathways in canine venous endothelial cells. Biochem J 297 1994209215
    Crossref | PubMed | ISI | Google Scholar
  • 151 Donnell-Tormey J, Nathan CF, Lanks K, DeBoer CJ, de la Harpe J Secretion of pyruvate. An antioxidant defense of mammalian cells. J Exp Med 165 1987500514
    Crossref | PubMed | ISI | Google Scholar
  • 152 Dosquet C, Weill D, Wautier JL Molecular mechanism of blood monocyte adhesion to vascular endothelial cells. Nouv Rev Fr Hematol 34 Suppl 1992S55S59
    PubMed | Google Scholar
  • 153 Downey JM Free radicals and their involvement during long-term myocardial ischemia and reperfusion. Annu Rev Physiol 52 1990487504
    Crossref | PubMed | ISI | Google Scholar
  • 154 Dreher D, Junod AF Differential effects of superoxide, hydrogen peroxide, and hydroxyl radical on intracellular calcium in human endothelial cells. J Cell Physiol 162 1995147153
    Crossref | PubMed | ISI | Google Scholar
  • 155 Dreher D, Junod AF Role of oxygen free radicals in cancer development. Eur J Cancer 32A 19963038
    Crossref | PubMed | ISI | Google Scholar
  • 156 Dröge W, Breitkreutz R N-acetyl-cysteine in the therapy of HIV-positive patients. Opin Clin Nutr Metab Care 2 1999493498
    Crossref | PubMed | Google Scholar
  • 157 Dröge W, Holm E Role of cysteine and glutathione in HIV infection and other diseases associated with muscle wasting and immunological dysfunction. FASEB J 11 199710771089
    Crossref | PubMed | ISI | Google Scholar
  • 158 Dröge W, Eck HP, Näher H, Pekar U, Daniel V Abnormal amino acid concentrations in the blood of patients with acquired immune deficiency syndrome (AIDS) may contribute to the immunological defect. Biol Chem Hoppe-Seyler 369 1988143148
    Crossref | PubMed | ISI | Google Scholar
  • 159 Dröge W, Mihm S, Bockstette M, Roth S Effect of reactive oxygen intermediates and antioxidants on proliferation and function of T lymphocytes. Methods Enzymol 234 1994135151
    Crossref | PubMed | ISI | Google Scholar
  • 160 Dröge W, Schulze-Osthoff K, Mihm S, Galter D, Schenk H, Eck HP, Roth S, Gmünder H Functions of glutathione and glutathione disulfide in immunology and immunopathology. FASEB J 8 199411311138
    Crossref | PubMed | ISI | Google Scholar
  • 161 Dukan S, Farewell A, Ballesteros M, Taddei F, Radman M, Nyström T Protein oxidation in response to increased transcriptional errors. Proc Natl Acad Sci USA 97 200057465749
    Crossref | PubMed | ISI | Google Scholar
  • 162 Dumont A, Hehner SP, Hofmann TG, Ueffing M, Dröge W, Schmitz ML Hydrogen peroxide-induced apoptosis is CD95-independent, requires the release of mitochondria-derived reactive oxygen species and the activation of NF-κB. Oncogene 18 1999747757
    Crossref | PubMed | ISI | Google Scholar
  • 163 Dupuy C, Ohayon R, Valent A, Noel-Hudson MS, Dème D, Virion A Purification of a novel flavoprotein involved in the thyroid NADPH oxidase. J Biol Chem 274 19993726537269
    Crossref | PubMed | ISI | Google Scholar
  • 164 Duthie GG, Robertson JD, Maughan RJ, Morrice PC Blood antioxidant status and erythrocyte lipid peroxidation following distance running. Arch Biochem Biophys 282 19907883
    Crossref | PubMed | ISI | Google Scholar
  • 165 Eck HP, Gmünder H, Hartmann M, Petzoldt D, Daniel V, Dröge W Low concentrations of acid soluble thiol (cysteine) in the blood plasma of HIV-1 infected patients. Biol Chem Hoppe-Seyler 370 1989101108
    Crossref | PubMed | ISI | Google Scholar
  • 166 Eck HP, Stahl-Hennig C, Hunsmann G, Dröge W Metabolic disorder as an early consequence of simian immunodeficiency virus infection in rhesus macaques. Lancet 338 1991346347
    Crossref | PubMed | ISI | Google Scholar
  • 167 Elbirt KK, Whitmarsh AJ, Davis RJ, Bonkovsky HL Mechanism of sodium arsenite-mediated induction of heme oxygenase-1 in hepatoma cells. Role of mitogen-activated protein kinases. J Biol Chem 273 199889228931
    Crossref | PubMed | ISI | Google Scholar
  • 168 Elroy-Stein O, Bernstein Y, Groner Y Overproduction of human Cn/Zn superoxide dismutase in transfected cells: extenuation of paraquat-mediated cytotoxicity and enhancement of lipid peroxidation. EMBO J 5 1986615622
    Crossref | PubMed | ISI | Google Scholar
  • 169 Endres R, Luz A, Schulze H, Neubauer H, Fütterer A, Holland SM, Wagner H, Pfeffer K Listeriosis in p47phox−/− and TRp55−/− mice: protection despite absence of ROI and susceptibility despite presence of RNI. Immunity 7 1997419432
    Crossref | PubMed | ISI | Google Scholar
  • 170 Estensen RD, Levy M, Klopp SJ, Galbraith AR, Mandel JS, Blomquist JA, Wattenberg LW N-acetylcysteine suppression of the proliferative index in the colon of patients with previous adenomatous colonic polyps. Cancer Lett 147 1999109114
    Crossref | PubMed | ISI | Google Scholar
  • 171 Esteve JM, Mompo J, de la Asuncion JG, Sastre J, Asensi M, Boix J, Vina JR, Vina J, Pallardó FV Oxidative damage to mitochondrial DNA and glutathione oxidation in apoptosis: studies in vivo and in vitro. FASEB J 13 199910551064
    Crossref | PubMed | ISI | Google Scholar
  • 172 Fahim MA, Robbins N Ultrastructural studies of young and old mouse neuromuscular junctions. J Neurocytol 11 1982641656
    Crossref | PubMed | Google Scholar
  • 173 Fandrey J, Frede S, Ehleben W, Porwol T, Acker H, Jelkmann W Cobalt chloride and desferrioxamine antagonize the inhibition of erythropoietin production by reactive oxygen species. Kidney Int 51 1997492496
    Crossref | PubMed | ISI | Google Scholar
  • 174 Fandrey J, Frede S, Jelkmann W Role of hydrogen peroxide in hypoxia-induced erythropoietin production. Biochem J 303 1994507510
    Crossref | PubMed | ISI | Google Scholar
  • 175 Fantus G, Kadota S, Deragon G, Foster B, Posner BI Pervanadate [peroxide(s) of vanadate] mimics insulin action in rat adipocytes via activation of the insulin receptor tyrosine kinase. Biochemistry 28 198988648871
    Crossref | PubMed | ISI | Google Scholar
  • 176 Fart SB, Kogoma T Oxidative stress responses in Escherichia coli and Salmonella typhimurium. Microbiol Rev 55 1991561585
    PubMed | Google Scholar
  • 177 Feaster WW, Kwok LW, Epstein CJ Dosage effects for superoxide dismutase-1 on nucleated cells aneuploid for chromosome 21. Am J Hum Genet 29 1977563570
    PubMed | ISI | Google Scholar
  • 178 Feng Xia LY, Garcia GE, Hwang D, Wilson CB Involvement of reactive oxygen intermediates in cyclooxygenase-2 expression induced by interleukin-1, tumor necrosis factor-alpha, and lipopolysaccharide. J Clin Invest 95 199516691675
    Crossref | PubMed | ISI | Google Scholar
  • 179 Firestein GS, Echeverri F, Yeo M, Zvaifler NJ, Green DR Somatic mutations in the p53 tumor suppressor gene in rheumatoid arthritis synovium. Proc Natl Acad Sci USA 94 19971089510900
    Crossref | PubMed | ISI | Google Scholar
  • 180 Flescher E, Ledbetter JA, Schieven GL, Vela-Roch N, Fossum D, Dang H, Ogawa N, Talal N Longitudinal exposure of human T lymphocytes to weak oxidative stress suppresses transmembrane and nuclear signal transduction. J Immunol 153 199448804889
    PubMed | ISI | Google Scholar
  • 181 Flier JS, Underhill LH Caloric intake and aging. N Engl J Med 337 1997986994
    Crossref | PubMed | ISI | Google Scholar
  • 182 Flint AJ, Tiganis T, Barford D, Tonks NK Development of "substrate trapping" mutants to identify physiological substrates of protein tyrosine phosphatases. Proc Natl Acad Sci USA 94 199716801685
    Crossref | PubMed | ISI | Google Scholar
  • 183 Flohe L, Brigelius R, Saliou C, Traber MG, Packer L Redox regulation of NF-κB activation. Free Radical Biol Med 22 199711151126
    Crossref | PubMed | ISI | Google Scholar
  • 184 Flores-Riveros JR, Sibley E, Kastelic T, Lane MD Substrate phosphorylation catalyzed by the insulin receptor tyrosine kinase. J Biol Chem 264 19892155721572
    PubMed | ISI | Google Scholar
  • 185 Force T, Bonventre JV, Heidecker G, Rapp U, Avruch J, Kyriakis JM Enzymatic characteristics of the c-Raf-1 protein kinase. Proc Natl Acad Sci USA 91 199412701274
    Crossref | PubMed | ISI | Google Scholar
  • 186 Frenette PS, Wagner DD Adhesion molecules. Part I. N Engl J Med 334 199615261529
    Crossref | PubMed | ISI | Google Scholar
  • 187 Fridovich I The biology of oxygen radicals. Science 201 1978875880
    Crossref | PubMed | ISI | Google Scholar
  • 188 Friedl HP, Till GO, Ryan US, Ward PA Mediator-induced activation of xanthine oxidase in endothelial cells. FASEB J 3 198925122518
    Crossref | PubMed | ISI | Google Scholar
  • 189 Fuchs SY, Adler V, Pincus MR, Ronai Z MEKK1/JNK signaling stabilizes and activates p53. Proc Natl Acad Sci USA 95 19981054110546
    Crossref | PubMed | ISI | Google Scholar
  • 190 Furchgott RF Interaction of H2O2 and NO in modifying tone in vascular smooth muscle: the SOD paradox. Resistance Arteries, Structure and Function , Mulvany MJ 1991216220ElsevierNew York
    Google Scholar
  • 191 Furukawa T, Itoh M, Krueger NX, Streuli M, Saito H Specific interaction of the CD45 protein-tyrosine phosphatase with tyrosine-phosphorylated CD3 zeta chain. Proc Natl Acad Sci USA 91 19941092810932
    Crossref | PubMed | ISI | Google Scholar
  • 192 Galter D, Mihm S, Dröge W Distinct effects of glutathione disulphide on the nuclear transcription factor kappa B and the activator protein-1. Eur J Biochem 221 1994639648
    Crossref | PubMed | Google Scholar
  • 193 Garbe A, Buck J, Hämmerling U Retinoids are important cofactors in T cell activation. J Exp Med 176 1992109117
    Crossref | PubMed | ISI | Google Scholar
  • 194 Garcia JH, Lassen NA, Weiller C, Sperling B, Nakagawara J Ischemic stroke and incomplete infarction. Stroke 27 1996761765
    Crossref | PubMed | ISI | Google Scholar
  • 195 Gaudu P, Weiss B SoxR, a [2Fe-2S] transcription factor, is active only in its oxidized form. Proc Natl Acad Sci USA 93 19961009410098
    Crossref | PubMed | ISI | Google Scholar
  • 196 Gersh BJ Current issues in reperfusion therapy. Am J Cardiol 82 19983P11P
    Crossref | PubMed | ISI | Google Scholar
  • 197 Gerzer R, Bohme E, Hofmann F, Schultz G Soluble guanylate purified from bovine lung contains heme and copper. FEBS Lett 132 19817174
    Crossref | PubMed | ISI | Google Scholar
  • 198 Ghosh S, Gifford AM, Rivière LR, Tempst P, Nolan GP, Baltimore D Cloning of the p50 DNA binding subunit of NFκB: homology to rel and dorsal. Cell 62 199010191029
    Crossref | PubMed | ISI | Google Scholar
  • 199 Gmünder H, Eck H-P, Benninghoff B, Roth S, Dröge W Macrophages regulate intracellular glutathione levels of lymphocytes. Evidence for an immunoregulatory role of cysteine. Cell Immunol 129 19903246
    Crossref | PubMed | ISI | Google Scholar
  • 200 Godon C, Lagniel G, Lee J, Buhler JM, Kieffer S, Perrot M, Boucherie H, Toledano MB, Labarre J The H2O2 stimulon in Saccharomyces cerevisiae. J Biol Chem 273 19982248022489
    Crossref | PubMed | ISI | Google Scholar
  • 201 Gohil K, Viguie C, Stanley WC, Brooks GA, Packer L Blood glutathione oxidation during human exercise. J Appl Physiol 64 1988115119
    Link | ISI | Google Scholar
  • 202 Goldspink DF The influence of immobilization and stretch on protein turnover of rat skeletal muscle. J Physiol (Lond) 264 1977267282
    Crossref | Google Scholar
  • 203 Goode HF, Webster NR, Howdle PD, Leek JP, Lodge JP, Sadek SA, Walker BE Reperfusion injury, antioxidants and hemodynamics during orthotopic liver transplantation. Hepatology 19 1994354359
    Crossref | PubMed | ISI | Google Scholar
  • 204 Gopalarishna R, Anderson WB Ca2+- and phospholipid-independent activation of protein kinase C by selective oxidative modification of the regulatory domain. Proc Natl Acad Sci USA 86 198967586762
    Crossref | PubMed | ISI | Google Scholar
  • 205 Gotoh Y, Cooper JA Reactive oxygen species- and dimerization-induced activation of apoptosis signal-regulating kinase 1 in tumor necrosis factor-α signal transduction. J Biol Chem 273 19981747717482
    Crossref | PubMed | ISI | Google Scholar
  • 206 Goulet JL, Snouwaert JN, Latour AM, Coffman TM, Koller BH Altered inflammatory responses in leukotriene-deficient mice. Proc Natl Acad Sci USA 91 19941285212856
    Crossref | PubMed | ISI | Google Scholar
  • 207 Gow AJ, Stamler JS Reactions between nitric oxide and haemoglobin under physiological conditions. Nature 391 1998169173
    Crossref | PubMed | ISI | Google Scholar
  • 208 Granger DN Role of xanthine oxidase and granulocytes in ischemia-reperfusion injury. Am J Physiol Heart Circ Physiol 255 1988H1269H1275
    Link | Google Scholar
  • 209 Greiner EF, Guppy M, Brand K Glucose is essential for proliferation and the glycolytic enzyme induction that provokes a transition to glycolytic energy production. J Biol Chem 269 19943148431490
    PubMed | ISI | Google Scholar
  • 210 Griendling KK, Minieri CA, Ollerenshaw JD, Alexander RW Angiotensin II stimulates NADH and NADPH oxidase activity in cultured vascular smooth muscle cells. Circ Res 74 199411411148
    Crossref | PubMed | ISI | Google Scholar
  • 211 Griendling KK, Sorescu D, Lassègue B, Ushio-Fukai M Modulation of protein kinase activity and gene expression by reactive oxygen species and their role in vascular physiology and pathophysiology. Arterioscler Thromb Vasc Biol 20 200021752183
    Crossref | PubMed | ISI | Google Scholar
  • 212 Griendling KK, Sorescu D, Ushio-Fukai M NAD(P)H oxidase. Role in cardiovascular biology and disease. Circ Res 86 2000494501
    Crossref | PubMed | ISI | Google Scholar
  • 213 Griffith CE, Zhang W, Wange RL ZAP-70-dependent and -independent activation of Erk in Jurkat T cells. J Biol Chem 273 19981077110776
    Crossref | PubMed | ISI | Google Scholar
  • 214 Gringhuis SI, Leow A, Papendrecht-van der Voort EAM, Remans PHJ, Breedveld FC, Verweij CL Displacement of linker for activation of T cells from the plasma membrane due to redox balance alterations results in hyporesponsiveness of synovial fluid T lymphocytes in rheumatoid arthritis. J Immunol 164 200021702179
    Crossref | PubMed | ISI | Google Scholar
  • 215 Griscavage JM, Fukuto JM, Komori Y, Ignarro LJ Nitric oxide inhibits neuronal nitric oxide synthase by interacting with their heme prosthetic group: role of tetrahydrobiopterin in modulating inhibitory action of nitric oxide. J Biol Chem 269 19942164421649
    PubMed | ISI | Google Scholar
  • 216 Groß A, Hack V, Stahl-Hennig C, Dröge W Elevated hepatic γ-glutamylcysteine synthetase activity and abnormal sulfate levels in liver and muscle tissue may explain abnormal cysteine and glutathione levels in SIV-infected rhesus macaques. AIDS Res Hum Retroviruses 12 199616391641
    Crossref | PubMed | ISI | Google Scholar
  • 217 Groß S, Knebel A, Tenev T, Neininger A, Gaestel M, Herrlich P, Böhmer FD Inactivation of protein tyrosine phosphatases as mechanism of UV-induced signal transduction. J Biol Chem 274 19992637826386
    Crossref | PubMed | ISI | Google Scholar
  • 218 Grubeck-Loebenstein B, Lechner H, Trieb K Long-term in vitro growth of human T cell clones: can postmitotic "senescent" cell populations be defined? Int Arch Allergy Immunol 104 1994232239
    Crossref | PubMed | ISI | Google Scholar
  • 219 Grune T, Reinheckel T, Davies KJA Degradation of oxidized proteins in K562 human hematopoietic cells by proteasome. J Biol Chem 271 19961550415509
    Crossref | PubMed | ISI | Google Scholar
  • 220 Grune T, Reinheckel T, Davies KJA Degradation of oxidized proteins in mammalian cells. FASEB J 11 1997526534
    Crossref | PubMed | ISI | Google Scholar
  • 221 Gulbis JM, Mann S, MacKinnon R Structure of a voltage-dependent K+ channel β subunit. Cell 97 1999943952
    Crossref | PubMed | ISI | Google Scholar
  • 222 Guyton KZ, Liu Y, Gorospe M, Xu Q, Holbrook NJ Activation of mitogen-activated protein kinase by H2O2. Role in cell survival following oxidant injury. J Biol Chem 271 199641384142
    Crossref | PubMed | ISI | Google Scholar
  • 223 Ha HC, Thiagalingam A, Nelkin BD, Casero RA Reactive oxygen species are critical for the growth and differentiation of medullary thyroid carcinoma cells. Clin Cancer Res 6 200037833787
    PubMed | ISI | Google Scholar
  • 224 Hack V, Breitkreutz R, Kinscherf R, Röhrer H, Bärtsch P, Taut F, Benner A, Dröge W The redox state as a correlate of senescence and wasting and as a target for therapeutic intervention. Blood 92 19985967
    PubMed | ISI | Google Scholar
  • 225 Hack V, Schmid D, Breitkreutz R, Stahl-Hennig C, Drings P, Kinscherf R, Taut F, Holm E, Dröge W Cystine levels, cystine flux and protein catabolism in cancer cachexia, HIV/SIV infection and senescence. FASEB J 11 19978492
    Crossref | PubMed | ISI | Google Scholar
  • 226 Hallbrucker C, Ritter M, Lang F, Gerok W, Häussinger D Hydroperoxide metabolism in rat liver. K+ channel activation, cell volume changes and eicosanoid formation. Eur J Biochem 211 1993449458
    Crossref | PubMed | Google Scholar
  • 227 Halliwell B Free radicals, reactive oxygen species and human disease: a critical evaluation with special reference to atherosclerosis. Br J Exp Pathol 70 1989737757
    PubMed | Google Scholar
  • 228 Halliwell B, Gutteridge JMC Free Radicals in Biology and Medicine (2nd ed.). 1989ClarendonOxford, UK
    Google Scholar
  • 229 Hampton MB, Kettle AJ, Winterbourn CC Inside the neutrophil phagosome: oxidants, myeloperoxidase, and bacterial killing. Blood 92 199830073017
    PubMed | ISI | Google Scholar
  • 230 Hamuro J, Murata Y, Suzuki M, Takatsuki F, Suga T The triggering and healing of tumor stromal inflammatory reactions regulated by oxidative and reductive macrophages. Gann Monograph Cancer Res 48 1999153164
    Google Scholar
  • 231 Hansen LL, Ikeda Y, Olsen GS, Busch AK, Mosthaf L Insulin signaling is inhibited by micromolar concentrations of H2O2. J Biol Chem 274 19992507825084
    Crossref | PubMed | ISI | Google Scholar
  • 232 Hardwick JS, Sefton BM Activation of the Lck tyrosine protein kinase by hydrogen peroxide requires the phosphorylation of Tyr-394. Proc Natl Acad Sci USA 92 199545274531
    Crossref | PubMed | ISI | Google Scholar
  • 233 Harley CB, Futcher AB, Greider CW Telomeres shorten during ageing of human fibroblasts. Nature 345 1990458460
    Crossref | PubMed | ISI | Google Scholar
  • 234 Harman D Aging: a theory based on free radical and radiation chemistry. J Gerontol 11 1956298300
    Crossref | PubMed | Google Scholar
  • 235 Harman D The aging process. Proc Natl Acad Sci USA 78 198171247128
    Crossref | PubMed | ISI | Google Scholar
  • 236 Harper JW, Elledge SJ, Keyomarsi K, Dynlacht B, Tsai L, Zhang P, Dobrowolski S, Bai C, Connel-Crowley L, Swindell E, Fox MP, Wei N Inhibition of cyclin-dependent kinases by p21. Mol Biol Cell 6 1995387400
    Crossref | PubMed | ISI | Google Scholar
  • 237 Hartsfield CL, Alam J, Cook JL, Choi AM Regulation of heme oxygenase-1 gene expression in vascular smooth muscle cells by nitric oxide. Am J Physiol Lung Cell Mol Physiol 273 1997L980L988
    Link | ISI | Google Scholar
  • 238 Hassan HM, Fridovich I Regulation of the synthesis of superoxide dismutase in Escherichia coli. J Biol Chem 252 197776677672
    PubMed | ISI | Google Scholar
  • 239 Hayashi T, Ueono Y, Okamoto T Oxidoreductive regulation of nuclear factor κB. Involvement of a cellular reducing catalyst thioredoxin. J Biol Chem 268 19931138011388
    PubMed | ISI | Google Scholar
  • 240 Hayes GR, Lockwood DH Role of insulin receptor phosphorylation in the insulinomimetic effects of hydrogen peroxide. Proc Natl Acad Sci USA 84 198781158119
    Crossref | PubMed | ISI | Google Scholar
  • 241 Hayflick L The limited in vitro lifetime of human diploid cell strains. Exp Cell Res 37 1965614636
    Crossref | PubMed | ISI | Google Scholar
  • 242 Hayflick L Theories of biological aging. Exp Gerontol 20 1985145159
    Crossref | PubMed | ISI | Google Scholar
  • 243 Hayflick L, Moorhead P The serial cultivation of human diploid cell strains. Exp Cell Res 25 1961585621
    Crossref | PubMed | ISI | Google Scholar
  • 244 He J, Kryger MH, Zorick FJ, Conway W, Roth T Mortality and apnea index in obstructive sleep apnea (experience in 385 male patients). Chest 94 1988914
    Crossref | PubMed | ISI | Google Scholar
  • 245 Hecht D, Zick Y Selective inhibition of protein tyrosine phosphatase activities by H2O2 and vanadate in vitro. Biochem Biophys Res Commun 188 1992773779
    Crossref | PubMed | ISI | Google Scholar
  • 246 Heeneman S, Haendeler J, Saito Y, Ishida M, Berk BC Angiotensin II induces transactivation of two different populations of the platelet-derived growth factor β receptor. J Biol Chem 275 20001592615932
    Crossref | PubMed | ISI | Google Scholar
  • 247 Heffetz D, Bushkin I, Dror R, Zick Y The insulinomimetic agents H2O2 and vanadate stimulate protein tyrosine phosphorylation in intact cells. J Biol Chem 265 199028962902
    PubMed | ISI | Google Scholar
  • 248 Hehner SP, Breitkreutz R, Shubinsky G, Unsoeld H, Schulze-Osthoff K, Schmitz ML, Dröge W Enhancement of T cell receptor signaling by a mild oxidative shift in the intracellular thiol pool. J Immunol 165 200043194328
    Crossref | PubMed | ISI | Google Scholar
  • 249 Hennet T, Richter C, Peterhans E Tumor necrosis factor-α induces superoxide anion generation in mitochondria of L929 cells. Biochem J 289 1993587592
    Crossref | PubMed | ISI | Google Scholar
  • 250 Henry Y, Lepoivre M, Drapier JC, Ducrocq C, Boucher JL, Guissani A EPR characterization of molecular targets for NO in mammalian cells and organelles. FASEB J 7 199311241134
    Crossref | PubMed | ISI | Google Scholar
  • 251 Herrlich P, Böhmer FD Redox regulation of signal transduction in mammalian cells. Biochem Pharmacol 59 20003541
    Crossref | PubMed | ISI | Google Scholar
  • 252 Herrmann G, Brenneisen P, Wlaschek M, Wenk J, Faisst K, Quel G, Hommel C, Goerz G, Ruzicka T, Krieg T, Sies H, Scharffetter-Kochanek K Psoralen photoactivation promotes morphological and functional changes in fibroblasts in vitro reminiscent of cellular senescence. J Cell Sci 111 1998759767
    PubMed | ISI | Google Scholar
  • 253 Heunks LM, Vina J, Herwaarden CL, Folgering HT, Gimeno A, Dekhuijzen PN Xanthine oxidase is involved in exercise-induced oxidative stress in chronic obstructive pulmonary disease. Am J Physiol Regulatory Integrative Comp Physiol 277 1999R1697R1704
    Link | ISI | Google Scholar
  • 254 Heymsfield SB, Clifford B, McManus CB Tissue of weight loss in cancer patients. Cancer 55 1985238249
    Crossref | PubMed | ISI | Google Scholar
  • 255 Hildalgo E, Bollinger JM, Bradley TM, Walsh CT, Demple B Binuclear [2Fe-2S] clusters in the Escherichia coli SoxR protein and role of the metal centers in transcription. J Biol Chem 270 19952090820914
    Crossref | PubMed | ISI | Google Scholar
  • 256 Hjeltnes N, Galuska D, Björnholm M, Aksnes AK, Lannem A, Zierath JR, Wallberg-Henrisson H Exercise-induced overexpression of key regulatory proteins involved in glucose uptake and metabolism in tetraplegic persons: molecular mechanism for improved glucose homeostasis. FASEB J 12 199817011712
    Crossref | PubMed | ISI | Google Scholar
  • 257 Hockenbery DM, Oltvai ZN, Yin XM, Milliman CL, Korsmeyer SJ Bcl-2 functions in an antioxidant pathway to prevent apoptosis. Cell 75 1993241251
    Crossref | PubMed | ISI | Google Scholar
  • 258 Holliday R Towards a biological understanding of the ageing process. Perspect Biol Med 32 1988109123
    Crossref | PubMed | ISI | Google Scholar
  • 259 Holloszy JO, Booth FW Biochemical adaptations to endurance exercise in muscle. Annu Rev Physiol 38 1976273291
    Crossref | PubMed | ISI | Google Scholar
  • 260 Honda Y, Honda S The daf-2 gene network for longevity regulates oxidative stress resistance and Mn-superoxide dismutase gene expression in Caenorhabditis elegans. FASEB J 13 199913851393
    Crossref | PubMed | ISI | Google Scholar
  • 261 Hortin GL, Landt M, Powderly WG Changes in plasma amino acid concentrations in response to HIV-1 infection. Clin Chem 40 1994785789
    Crossref | PubMed | ISI | Google Scholar
  • 262 Houston ME Adaptions in skeletal muscle to training and detraining: the role of protein synthesis and degradation. Biochemistry of Exercise VI , Saltin B 16 19866374Human KineticsChampaign, IL (Int Ser Sports Sci)
    Google Scholar
  • 263 Hoyos B, Imam A, Chua R, Swenson C, Tong GX, Levi E, Noy N, and Hämmerling U. The cysteine-rich regions of the regulatory domains of Raf and protein kinase C as retinoid receptors. J Exp Med. In press.
    Google Scholar
  • 264 Huang LE, Arany Z, Livingston DM, Bunn FH Activation of hypoxia-inducible transcription factor depends primarily upon redox-sensitive stabilization of its alpha subunit. J Biol Chem 271 19963225332259
    Crossref | PubMed | ISI | Google Scholar
  • 265 Huang LE, Gu J, Schau M, Bunn HF Regulation of hypoxia-inducible factor 1α is mediated by it oxygen-dependent degradation domain via the ubiquitin-proteasome pathway. Proc Natl Acad Sci USA 95 199879877992
    Crossref | PubMed | ISI | Google Scholar
  • 266 Hubbard SR Crystal structure of the activated insulin receptor tyrosine kinase in complex with peptide substrate and ATP analog. EMBO J 16 199755725581
    Crossref | PubMed | ISI | Google Scholar
  • 267 Hubbard SR, Wei L, Ellis L, Hendrickson WA Crystal structure of the tyrosine kinase domain of the human insulin receptor. Nature 372 1994746754
    Crossref | PubMed | ISI | Google Scholar
  • 268 Hug H, Enari M, Nagata S No requirement of reactive oxygen intermdiates in Fas-mediated apoptosis. FEBS Lett 351 1994311313
    Crossref | PubMed | ISI | Google Scholar
  • 269 Hussain SP, Aguilar F, Amstad P, Cerutti P Oxy-radical induced mutagenesis of hotspot codons 248 and 249 of the human p53 gene. Oncogene 9 199422772281
    PubMed | ISI | Google Scholar
  • 270 Ichijo H, Nishida E, Irie K, Dijke P, Saitoh M, Moriguchi T, Takagi M, Matsumoto K, Miyzaono K, Gotoh Y Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science 275 19979094
    Crossref | PubMed | ISI | Google Scholar
  • 271 Ido Y, Kilo C, Williamson JR Cytosolic NADH/NAD+ free radicals and vascular dysfunction in early diabetes mellitus. Diabetologia 40 1997115117
    Crossref | PubMed | ISI | Google Scholar
  • 272 Ignarro LJ, Kadowitz PJ The pharmacological and physiological role of cyclic GMP in vascular smooth muscle relaxation. Ann Pharmacol Toxicol 25 1985171191
    Crossref | PubMed | ISI | Google Scholar
  • 273 Ignarro LJ, Wood KS, Wolin MS Regulation of purified soluble guanylate cyclase by porphyrins and metalloporphrins: a unifying concept. Adv Cyclic Nucleotide Protein Phosphorylation Res 17 1984267274
    PubMed | Google Scholar
  • 274 Imam A, Hoyos B, Swenson C, Levi E, Chua R, Viriya E, Hämmerling U Retinoids as ligands and coactivators of protein kinase C alpha. FASEB J 15 20012830
    Crossref | PubMed | ISI | Google Scholar
  • 275 Ip YT, Davis RJ Signal transduction by the c-Jun N-terminal kinase (JNK)-from inflammation to development. Curr Opin Cell Biol 10 1998205219
    Crossref | PubMed | ISI | Google Scholar
  • 276 Irani K, Xia Y, Zweier JL, Sollott SJ, Der CJ, Fearon ER, Sundaresan M, Finkel T, Goldschmidt-Clermont PJ Mitogenic signaling mediated by oxidants in Ras-transformed fibroblasts. Science 275 199716491652
    Crossref | PubMed | ISI | Google Scholar
  • 277 Ishii N Oxidative stress and aging in Caenorhabditis elegans. Free Radical Res 33 2000857864
    Crossref | PubMed | ISI | Google Scholar
  • 278 Ishii T, Itoh K, Sato H, Bannai S Oxidative stress-inducible proteins in macrophages. Free Radical Res 31 1999351355
    Crossref | PubMed | ISI | Google Scholar
  • 279 Jabs T, Dietrich RA, Dangl JL Initiation of runaway cell death in an Arabidopsis mutant by extracellular superoxide. Science 273 199618531856
    Crossref | PubMed | ISI | Google Scholar
  • 280 Jackson MJ, Edwards RHT, Symons MCR Electron spin resonance studies of intact mammalian skeletal muscle. Biochim Biophys Acta 847 1985185190
    Crossref | PubMed | ISI | Google Scholar
  • 281 Jacobson MD, Burne JF, Raff MC Programmed cell death and Bcl-2 protection in the absence of a nucleus. EMBO J 13 199418991910
    Crossref | PubMed | ISI | Google Scholar
  • 282 Jaeschke H, Smith CV, Mitchell JR Hypoxic damage generates reactive oxygen species in isolated perfused rat liver. Biochem Biophys Res Commun 150 1988568574
    Crossref | PubMed | ISI | Google Scholar
  • 283 Janssen YM, Matalon S, Mossman BT Differential induction of c-fos, c-jun, and apoptosis in lung epithelial cells exposed to ROS or RNS. Am J Physiol Lung Cell Mol Physiol 273 1997L789L796
    Link | ISI | Google Scholar
  • 284 Ji LL, Dillon D, Wu E Alteration of antioxidant enzymes with aging in skeletal muscle and liver. Am J Physiol Regulatory Integrative Comp Physiol 258 1990R918R923
    Link | ISI | Google Scholar
  • 285 Jia Z, Barford D, Flint AJ, Tonks NK Structural basis for phosphotyrosine peptide recognition by protein tyrosine phosphatase 1B. Science 268 199517541758
    Crossref | PubMed | ISI | Google Scholar
  • 286 Jocelyn PC Biochemistry of the SH Group. 197271AcademicLondon
    Google Scholar
  • 287 Johnson TM, Yu ZX, Ferrans VJ, Lowenstein RA, Finkel T Reactive oxygen species are downstream mediators of p53-dependent apoptosis. Proc Natl Acad Sci USA 93 19961184811852
    Crossref | PubMed | ISI | Google Scholar
  • 288 Jones SA, Hancock J, Jones OTG, Neubauer A, Topley N The expression of NADPH oxidase components in human glomerular mesangial cells: detection of protein and mRNA for p47phax, p67phax, and p22phax. J Am Soc Nephrol 5 199514831491
    PubMed | ISI | Google Scholar
  • 289 Jones SA, O'Donnell VB, Wood JD, Broughton JP, Hughes EJ, Jones OT Expression of phagocyte NADPH oxidase components in human endothelial cells. Am J Physiol Cell Physiol 271 1996C626C634
    ISI | Google Scholar
  • 290 Jones SA, Wood JD, Coffey MJ, Jones OT The functional expression of p47-phox and p67-phox may contribute to the generation of superoxide by an NADPH oxidase-like system in human fibroblasts. FEBS Lett 355 1994178182
    Crossref | PubMed | ISI | Google Scholar
  • 291 Jungermann K, Kietzmann T Role of oxygen in the zonation of carbohydrate metabolism and gene expression in liver. Kidney Int 51 1997402412
    Crossref | PubMed | ISI | Google Scholar
  • 292 Kahn CR Insulin action, diabetogenes, and the cause of type II diabetes (Banting lecture). Diabetes 43 199410661084
    Crossref | PubMed | ISI | Google Scholar
  • 293 Kaltschmidt B, Uherek M, Volk B, Baeuerle PA, Katlschmidt C Transcription factor NF-κB is activated in primary neurons by amyloid beta peptides and in neurons surrounding early plaques from patients with Alzheimer disease. Proc Natl Acad Sci USA 94 199726422647
    Crossref | PubMed | ISI | Google Scholar
  • 294 Kamata H, Hirata H Redox regulation of cellular signalling. Cell Signal 11 1999114
    Crossref | PubMed | ISI | Google Scholar
  • 295 Karin M The regulation of AP-1 activity by mitogen-activated protein kinases. J Biol Chem 270 19951648316486
    Crossref | PubMed | ISI | Google Scholar
  • 296 Karin M, Liu Z, Zandi E AP-1 function and regulation. Curr Opin Cell Biol 9 1997240246
    Crossref | PubMed | ISI | Google Scholar
  • 297 Karin M, Smeal T Control of transcription factors by signal transduction pathways: the beginning of the end. Trends Biochem Sci 17 1992418422
    Crossref | PubMed | ISI | Google Scholar
  • 298 Karpen CW, Pritchard KA, Arnold JH, Cornwell DG, Panganamala RV Restoration of prostacyclin/thromboxane A2 balance in the diabetic rat. Influence of dietary vitamin E. Diabetes 31 1982947951
    Crossref | ISI | Google Scholar
  • 299 Keisari Y, Braun L, Flescher E The oxidative burst and related phenomena in mouse macrophages elicited by different sterile inflammatory stimuli. Immunobiology 165 19837889
    Crossref | PubMed | ISI | Google Scholar
  • 300 Keyse SM, Tyrrell RM Heme oxygenase is the major 32-kDa stress protein induced in human skin fibroblasts by UVA radiation, hydrogen peroxide, and sodium arsenite. Proc Natl Acad Sci USA 86 198999103
    Crossref | PubMed | ISI | Google Scholar
  • 301 Kheradmand F, Werner E, Tremble P, Symons M, Werb Z Role of Rac1 and oxygen radicals in collagenase-1 expression induced by cell shape change. Science 280 1998898902
    Crossref | PubMed | ISI | Google Scholar
  • 302 Kieran M, Blank V, Logeat F, Vandekerckhove J, Lottspeich F, Le Bail O, Urban MB, Kourilsky P, Baeuerle PA, Israel A The DNA binding subunit of NFκB is identical to factor KBF1 and homologous to the rel oncogene product. Cell 62 199010071018
    Crossref | PubMed | ISI | Google Scholar
  • 303 Kiessling R, Grönberg A, Ivanyi J, Söderstrom K, Ferm M, Kleinau S, Nilsson E, Klareskog L Role of hsp 60 during autoimmune and bacterial inflammation. Immunol Rev 12 199191111
    Crossref | ISI | Google Scholar
  • 304 Kim YM, Talanian RV, Billiar TR Nitric oxide inhibits apoptosis by preventing increases in caspase-3-like activity via two distinct mechanisms. J Biol Chem 272 19973113831148
    Crossref | PubMed | ISI | Google Scholar
  • 305 Kinscherf R, Claus R, Wagner M, Gehrke C, Kamencic H, Hou D, Nauen O, Schmiedt W, Kovacs G, Pill J, Metz J, Deigner HP Apoptosis caused by oxidized LDL is manganese superoxide dismutase and p53 dependent. FASEB J 12 1998461467
    Crossref | PubMed | ISI | Google Scholar
  • 306 Kinscherf R, Deigner HP, Usinger C, Pill J, Wagner M, Kamencic H, Hou D, Schen M, Schmiedt W, Schrader M, Kovacs G, Kato K, Metz J Induction of mitochondrial manganese superoxide dismutase in macrophages by oxidized LDL: its relevance in atherosclerosis of humans and heritable hyperlipidemic rabbits. FASEB J 11 199713171328
    Crossref | PubMed | ISI | Google Scholar
  • 307 Kinscherf R, Fischbach T, Mihm S, Roth S, Hohenhaus-Sievert E, Weiss C, Edler L, PBärtsch, Dröge W Effect of glutathione depletion and oral N-acetyl-cysteine treatment on CD4+ and CD8+ cells. FASEB J 8 1994448451
    Crossref | PubMed | ISI | Google Scholar
  • 308 Kinscherf R, Kamencic H, Deigner HP, Pill J, Schmiedt W, Schrader M, Metz J Effect of alterations of blood cholesterol levels on macrophages in the myocardium of New Zealand White rabbits. J Leukoc Biol 62 1997719725
    Crossref | PubMed | ISI | Google Scholar
  • 309 Kinscherf R, Wagner M, Kamencic H, Bonaterra GA, Hou D, Schiele RA, Deigner HP, Metz J Characterization of apoptotic macrophages in atheromatous tissue of humans and heritable hyperlipidemic rabbits. Atherosclerosis 144 19993339
    Crossref | PubMed | ISI | Google Scholar
  • 310 Klebanoff SJ, Vadas MA, Harlan JM, Sparks LH, Gamble JR, Agosti JM, Waltersdorph AM Stimulation of neutrophils by tumor necrosis factor. J Immunol 136 198642204225
    PubMed | ISI | Google Scholar
  • 311 Klotz LO, Briviba K, Sies H Singlet oxygen mediates the activation of JNK by UVA radiation in human skin fibroblasts. FEBS Lett 408 1997289291
    Crossref | PubMed | ISI | Google Scholar
  • 312 Klotz LO, Pellieux C, Briviba K, Pierlot C, Aubry JM, Sies H Mitogen-activated protein kinase (p38-, JNK-, ERK-) activation pattern induced by extracellular and intracellular singlet oxygen and UVA. Eur J Biochem 260 1999917922
    Crossref | PubMed | Google Scholar
  • 313 Knebel A, Rahmsdorf HJ, Ullrich A, Herrlich P Dephosphorylation of receptor tyrosine kinases as target of regulation by radiation, oxidants or alkylating agents. EMBO J 15 199653145325
    Crossref | PubMed | ISI | Google Scholar
  • 314 Kondo H, Miura M, Itokawa Y Oxidative stress in skeletal muscle atrophied by immobilization. Acta Physiol Scand 142 1991527528
    Crossref | PubMed | Google Scholar
  • 315 Kondo H, Nishino K, Itokawa Y Hydroxyl radical generation in skletal muscle atrophied by immobilization. FEBS Lett 349 1994169172
    Crossref | PubMed | ISI | Google Scholar
  • 316 Konishi H, Tanaka M, Takemura Y, Matsuzaki H, Ono Y, Kikkawa U, Nishizuka Y Activation of protein kinase C by tyrosine phosphorylation in response to H2O2. Proc Natl Acad Sci USA 94 19971123311237
    Crossref | PubMed | ISI | Google Scholar
  • 317 Korsmeyer SJ Regulators of cell death. Trends Genet 11 1995101105
    Crossref | PubMed | ISI | Google Scholar
  • 318 Kovtun Y, Chiu WL, Tena G, Sheen J Functional analysis of oxidative stress-activated mitogen-activated protein kinase cascade in plants. Proc Natl Acad Sci USA 97 200029402945
    Crossref | PubMed | ISI | Google Scholar
  • 319 Kretz-Remy C, Bates EE, Arrigo AP Amino acid analogs activate NF-κB through redox-dependent IκB-α phosphorylation. Consequence on HIV-1 long terminal repeat activation. J Biol Chem 273 199831803191
    Crossref | PubMed | ISI | Google Scholar
  • 320 Kretz-Remy C, Mehlen P, Mirault ME, Arrigo AP Inhibition of IκBα phosphorylation and degradation and subsequent NF-κB activation by glutathione peroxidase overexpression. J Cell Biol 133 199610831093
    Crossref | PubMed | ISI | Google Scholar
  • 321 Krieger-Brauer H, Medda PK, Kather H Insulin-induced activation of NADPH-dependent H2O2 generation in human adipocyte plasma membranes is mediated by Gαi2. J Biol Chem 272 19971013510143
    Crossref | PubMed | ISI | Google Scholar
  • 322 Ksenzenko M, Konstantinov AA, Khomutov GB, Tikhonov AN, Ruuge EK Effect of electron transfer inhibitors on superoxide generation in the cytochrome bc1 site of the mitochondrial respiratory chain. FEBS Lett 155 19831924
    Crossref | PubMed | ISI | Google Scholar
  • 323 Kuge S, Jones N YAP-1 dependent activation of TRX2 is essential for the response of Saccharomyces cerevisiae to oxidative stress by hydroperoxides. EMBO J 13 1994655664
    Crossref | PubMed | ISI | Google Scholar
  • 324 Kuge S, Jones N, Nomoto A Regulation of yAP-1 nuclear localization in response to oxidative stress. EMBO J 16 199717101720
    Crossref | PubMed | ISI | Google Scholar
  • 325 Kukielka GL, Youker KA, Hawkins HK, Perrard JL, Michael LH, Ballantyne CM, Smith CW, Entman ML Regulation of ICAM-1 and IL-6 in myocardial ischemia: effect of reperfusion. Ann NY Acad Sci 723 1994258270
    Crossref | PubMed | ISI | Google Scholar
  • 326 Kumar S, Rabson AB, Gélinas C The RxxRxRxxC motif conserved in all Rel/κB proteins is essential for the DNA binding activity and redox regulation of the v-Rel oncoprotein. Mol Cell Biol 12 199230943106
    Crossref | PubMed | ISI | Google Scholar
  • 327 Kumasaka S, Shoji H, Okabe E Novel mechanisms involved in superoxide anion radical-triggered Ca2+ release from cardiac sarcoplasmic reticulum linked to cyclic ADP-ribose stimulation. Antiox Redox Signal 1 19995569
    Crossref | PubMed | ISI | Google Scholar
  • 328 Kunsch C, Medford RM Oxidative stress as a regulator of gene expression in the vasculature. Circ Res 85 1999753766
    Crossref | PubMed | ISI | Google Scholar
  • 329 Lahiri S, Acker H Redox-dependent binding of CO to heme protein controls Po 2-sensitive chemoreceptor discharge of the rat carotid body. Respir Physiol 115 1999169177
    Crossref | PubMed | Google Scholar
  • 330 Lamas S, Marsden PA, Li GK, Tempst P, Michel T Endothelial nitric oxide synthase: molecular cloning and characterization of a distinct constitutive enzyme isoform. Proc Natl Acad Sci USA 89 199263486352
    Crossref | PubMed | ISI | Google Scholar
  • 331 Lamb C, Dixon RA The oxidative burst in plant disease resistance. Annu Rev Plant Physiol Plant Mol Biol 48 1997251275
    Crossref | PubMed | ISI | Google Scholar
  • 332 Lamberts SWJ, van den Beld AW, van der Lely AJ The endocrinology of aging. Science 278 1997419424
    Crossref | PubMed | ISI | Google Scholar
  • 333 Lambeth JD, Cheng G, Arnold RS, Edens WA Novel homologs of gp91 phox. Trends Biochem Sci 25 2000459461
    Crossref | PubMed | ISI | Google Scholar
  • 334 Lander HM An essential role for free radicals and derived species in signal transduction. FASEB J 11 1997118124
    Crossref | PubMed | ISI | Google Scholar
  • 335 Lander HM, Hajjar DP, Hempstead BL, Mirza UA, Chait BT, Campbell S, Quilliam LA A molecular redox switch on p21(ras). Structural basis for the nitric oxide-p21(ras) interaction. J Biol Chem 272 199743234326
    Crossref | PubMed | ISI | Google Scholar
  • 336 Lander HM, Jacovina AT, Davis RJ, Tauras JM Differential activation of mitogen-induced protein kinases by nitric-oxide-related species. J Biol Chem 271 19961970519709
    Crossref | PubMed | ISI | Google Scholar
  • 337 Lander HM, Milbank AJ, Tauras JM, Hajjar DP, Hempstead BL, Schwartz GD, Kraemer RT, Mirza UA, Chait BT, Burk SC, Quilliam LA Redox regulation of cell signalling. Nature 381 1996380381
    Crossref | PubMed | ISI | Google Scholar
  • 338 Lander HM, Ogiste JS, Pearce SF, Levi R, Novogrodsky A Nitric oxide-stimulated guanine nucleotide exchange on p21ras. J Biol Chem 270 199570177020
    Crossref | PubMed | ISI | Google Scholar
  • 339 Lander HM, Sehajpal P, Levine DM, Novogrodsky A Activation of human peripheral blood mononuclear cells by nitric oxide-generating compounds. J Immunol 150 199315091516
    PubMed | ISI | Google Scholar
  • 340 Larsson R, Cerutti P Translocation and enhancement of phosphotransferase activity of protein kinase C following exposure in mouse epidermal cells to oxidants. Cancer Res 49 198956275632
    PubMed | ISI | Google Scholar
  • 341 Lazou A, Bogoyevitch MA, Clerk A, Fuller SJ, Marshall CJ, Sugden PH Regulation of mitogen-activated protein kinase cascade in adult rat heart preparations in vitro. Circ Res 75 1994932941
    Crossref | PubMed | ISI | Google Scholar
  • 342 Lee AC, Fenster BE, Ito H, Takeda K, Bae NS, Hirai T, Yu ZX, Ferrans VJ, Howard BH, Finkel T Ras proteins induce senescence by altering the intracellular levels of reactive oxygen species. J Biol Chem 274 199979367940
    Crossref | PubMed | ISI | Google Scholar
  • 343 Lee CK, Klopp RG, Weindruch R, Prolla TA Gene expression profile of aging and its retardation by caloric restriction. Science 285 199913901393
    Crossref | PubMed | ISI | Google Scholar
  • 344 Lee CK, Weindruch R, Prolla TA Gene expression profile of the aging brain in mice. Nature Genet 25 2000294297
    Crossref | PubMed | ISI | Google Scholar
  • 345 Lempicki RA, Kovacs JA, Baseler MW, Adelsberger JW, Dewar RL, Natarajan V, Bosche MC, Metcalf JA, Stevens RA, Lamberts LA, Alvord WG, Polis MA, Davey RT, Dimitrov DS, Lane HC Impact of HIV-1 infection and highly active antiretroviral therapy on the kinetics of CD4+ and CD8+ T cell turnover in HIV-infected patients. Proc Natl Acad Sci USA 97 20001377813783
    Crossref | PubMed | ISI | Google Scholar
  • 346 Leusen JHW, De KA, Hilarius PM, Ahlin A, Palmblad J, Smith CI, Diekmann D, Hall A, Verhoeven AJ, Roos D Disturbed interaction of p21-rac with mutated p67-phox causes chronic granulomatous disease. J Exp Med 184 199612431249
    Crossref | PubMed | ISI | Google Scholar
  • 347 Levine A, Tenhaken R, Dixon RA, Lamb C H2O2 from the oxidative burst orchestrates the plant hypersensitive disease resistance response. Cell 79 1994583593
    Crossref | PubMed | ISI | Google Scholar
  • 348 Lew H, Pyke S, Quintanilha A Changes in the glutathione status of plasma, liver and muscle following exhaustive excercise in rats. FEBS Lett 185 1985262266
    Crossref | PubMed | ISI | Google Scholar
  • 349 Li J, Billiar TR, Talanian RV, Kim YM Nitric oxide reversibility inhibits seven members of the caspase family via S-nitrosylation. Biochem Biophys Res Commun 240 1997419424
    Crossref | PubMed | ISI | Google Scholar
  • 350 Li N, Karin M Is NF-κB the sensor of oxidative stress? FASEB J 13 199911371143
    Crossref | PubMed | ISI | Google Scholar
  • 351 Lin YJ, Seroude L, Benzer S Extended life-span and stress resistance in the Drosophila mutant methuselah. Science 282 1998943946
    Crossref | PubMed | ISI | Google Scholar
  • 352 Linsley PS, Ledbetter JA The role of the CD28 receptor during T cell responses to antigen. Annu Rev Immunol 11 1993191212
    Crossref | PubMed | ISI | Google Scholar
  • 353 Liu H, Nishitoh H, Ichijo H, Kyriakis JM Activation of apoptosis signal-regulating kinase 1 (ASK1) by tumor necrosis factor receptor-associated factor 2 requires prior dissociation of the ASK1 inhibitor thioredoxin. Mol Cell Biol 20 200021982208
    Crossref | PubMed | ISI | Google Scholar
  • 354 Lo YY, Cruz TF Involvement of reactive oxygen species in cytokine and growth factor induction of c-fos expression in chondrocytes. J Biol Chem 270 19951172711730
    Crossref | PubMed | ISI | Google Scholar
  • 355 Lo YYC, Wong JMS, Cruz TF Reactive oxygen species mediate cytokine activation of c-Jun NH2-terminal kinases. J Biol Chem 271 19961570315707
    Crossref | PubMed | ISI | Google Scholar
  • 356 López-Barneo J, Pardal R, Montoro RJ, Smani T, Garcı́a-Hirschfeld J, Urena J K+ and Ca2+ channel activity and cytosolic [Ca2+] in oxygen-sensing tissues. Respir Physiol 115 1999215227
    Crossref | PubMed | Google Scholar
  • 357 Lopez-Collazo E, Mateo J, Miras-Portugal MT, Bosca L Requirement of nitric oxide and calcium mobilization for the induction of apoptosis in adrenal vascular endothelial cells. FEBS Lett 413 1997124128
    Crossref | PubMed | ISI | Google Scholar
  • 358 Los M, Dröge W, Stricker K, Baeuerle PA, Schulze-Osthoff K Hydrogen peroxide as a potent activator of T lymphocyte functions. Eur J Immunol 25 1995159165
    Crossref | PubMed | ISI | Google Scholar
  • 359 Los M, Schenk H, Hexel K, Baeuerle PA, Dröge W, Schulze-Osthoff K IL-2 gene expression and NF-κB activation through CD28 requires reactive oxygen production by 5-lipoxygenase. EMBO J 14 199537313740
    Crossref | PubMed | ISI | Google Scholar
  • 360 Loschen G, Azzi A, Flohe L Mitochondrial hydrogen peroxide formation. Alcohol and Aldehyde Metabolizing Systems , Thurman T, Yonetani T, Williams JR, Chance B 1974215229AcademicNew York
    Google Scholar
  • 361 Loschen G, Azzi A, Richter C, Flohe L Superoxide radicals as precursors of mitochondrial hydrogen peroxide. FEBS Lett 42 19746872
    Crossref | PubMed | ISI | Google Scholar
  • 362 Loschen G, Flohe L, Chance B Respiratory chain linked H2O2 production in pigeon heart mitochondria. FEBS Lett 18 1971261264
    Crossref | PubMed | ISI | Google Scholar
  • 363 Lovell MA, Ehmann WD, Mattson MP, Markesbery WR Elevated 4-hydroxynonenal in ventricular fluid in Alzheimer's disease. Neurobiol Aging 18 1997457461
    Crossref | PubMed | ISI | Google Scholar
  • 364 Low PA, Nickander KK Oxygen free radical effects in sciatic nerve in experimental diabetes. Diabetes 40 1991873877
    Crossref | PubMed | ISI | Google Scholar
  • 365 Luo Y, Roth GS The roles of dopamine oxidative stress and dopamine receptor signaling in aging and age-related neurodegeneration. Antioxidants Redox Signal 2 2000449460
    Crossref | PubMed | ISI | Google Scholar
  • 366 MacMicking J, Xie QW, Nathan C Nitric oxide and macrophage function. Annu Rev Immunol 15 1997323350
    Crossref | PubMed | ISI | Google Scholar
  • 367 Maki A, Berezesky IK, Fargnoli J, Holbrook NJ, Trump BF Role of Ca2+ in induction of c-fos, c-jun, and c-myc mRNA in rat PTE after oxidative stress. FASEB J 6 1992919924
    Crossref | PubMed | ISI | Google Scholar
  • 368 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 κB and activated protein-1. J Biol Chem 273 19981324513254
    Crossref | PubMed | ISI | Google Scholar
  • 369 Mans DRA, Lafleur MVM, Westmijize EJ, Horn IR, Bets D, Schuurhuis GJ, Lankelma J, Retel J Reactions of glutathione with the catechol, the ortho-quinone and the semi-quinone free radical of etoposide. Consequences for DNA inactivation. Biochem Pharmacol 43 199217611768
    Crossref | PubMed | ISI | Google Scholar
  • 370 Mapp PI, Grootveld MC, Blake DR Hypoxia, oxidative stress and rheumatoid arthritis. Br Med Bull 51 1995419436
    Crossref | PubMed | ISI | Google Scholar
  • 371 Marks PA, Bishop JS The glucose metabolism of patients with malignant disease and of normal subjects as studied by means of an intravenous glucose tolerance test. J Clin Invest 36 1957254264
    Crossref | PubMed | ISI | Google Scholar
  • 372 Matsuda Moriguchi ST, Koyasu S, Nishida E T-lymphocyte activation signals for interleukin-2 production involve activation of MKK6–p38 and MKK7-SAPK/JNK signaling pathways sensitive to cyclosporin A. J Biol Chem 273 199812378123782
    Crossref | PubMed | ISI | Google Scholar
  • 373 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 1996179185
    Crossref | PubMed | ISI | Google Scholar
  • 374 Matthews JR, Wakasugi N, Virelizier JL, Yodoi J, Hay RT Thioredoxin regulates the DNA binding activity of NF-κB by reduction of a disulphide bond involving cysteine 62. Nucleic Acids Res 20 199238213830
    Crossref | PubMed | ISI | Google Scholar
  • 375 Maulik N, Engelman RM, Rousou JA, Flack JE, Deaton DW, Das DK Ischemic preconditioning suppresses apoptosis by upregulating the antideath gene Bel-2. Surg Forum 49 1998209211
    Google Scholar
  • 376 Maulik N, Sato M, Price BD, Das D An essential role of NFκB in tyrosine kinase signaling of p38 MAP kinase regulation of myocardial adaptation to ischemia. FEBS Lett 429 1998365369
    Crossref | PubMed | ISI | Google Scholar
  • 377 Maulik N, Yoshida T, Engelman RM, Deaton DW, Flack JE, Rousou JA, Das DK Ischemic preconditioning attenuates apoptotic cell death associated with ischemia/reperfusion. Mol Cell Biochem 186 1998139145
    Crossref | PubMed | ISI | Google Scholar
  • 378 Maurice MM, Nakamura H, van der Voort EAM, van Vliet AI, Staal FJT, Tak PP, Breedveld FC, Verweij CL Evidence for the role of an altered redox state in hyporesponsiveness of synovial T cells in rheumatoid arthritis. J Immunol 158 199714581465
    PubMed | ISI | Google Scholar
  • 379 Maurice MM, van der Voort E, van Vliet AI, Tak PP, Breedveld FC, Verweij CL The rheumatoid joint: redox-paradox? Oxidative Stress in Cancer, AIDS and Neurodegenrative Diseases , Montagnier L, Olivier R, Pasquier C 1998517DekkerNew York
    Google Scholar
  • 380 May JM, DeHäen C Insulin-stimulated intracellular hydrogen peroxide production in rat epididymal fat cell. J Biol Chem 254 19792214
    PubMed | ISI | Google Scholar
  • 381 Maziere C, Auclair M, Rose-Robert F, Leflon P, Maziere JC Glucose-enriched medium enhances cell-mediated low density lipoprotein peroxidation. FEBS Lett 363 1995277279
    Crossref | PubMed | ISI | Google Scholar
  • 382 McArdle A, van der Meulem JH, Catapano M, Symons MCR, Faulkner JA, Jackson MJ Free radical activity during concentration-induced injury to the extensor digitorum longus muscle of rats. J Physiol (Lond) 487 1995157P158P
    Google Scholar
  • 383 McArdle A, van der Meulen JH, Catapano M, Symons MCR, Faulkner JA, Jackson MJ Free radical activity following contraction-induced injury to the extensor digitorum longus muscles of rats. Free Radical Biol Med 26 199910851091
    Crossref | PubMed | ISI | Google Scholar
  • 384 McCarron M, Osborne Y, Stroy CJ, Dempsey IL, Tuner DR, Morley AA Effect of age on lymphocyte proliferation. Mech Ageing Dev 41 1987211218
    Crossref | PubMed | ISI | Google Scholar
  • 385 McCord JM Oxygen-derived free radicals in postischemic tissue injury. N Engl J Med 312 1985159163
    Crossref | PubMed | ISI | Google Scholar
  • 386 McCord JM, Fridovich I Superoxide dismutase: an enzymic function for erythrocuprein (hemocuprein). J Biol Chem 244 196960496055
    PubMed | ISI | Google Scholar
  • 387 Meier B, Radeke HH, Selle S, Younes M, Sies H, Resch K, Habermehl GG Human fibroblasts release reactive oxygen species in response to interleukin-1 or tumor necrosis factor-α. Biochem J 263 1989539545
    Crossref | PubMed | ISI | Google Scholar
  • 388 Melov S, Ravenscroft J, Malik S, Gill MS, Walker DW, Clayton PE, Wallace DC, Malfroy B, Doctrow SR, Lithgow GJ Extension of life-span with superoxide dismutase/catalase mimetics. Science 289 200015671569
    Crossref | PubMed | ISI | Google Scholar
  • 389 Meyer M, Schreck R, Baeuerle PA H2O2 and antioxidants have opposite effects on activation of NF-kB and AP-1 in intact cells: AP-1 as secondary antioxidant response factor. EMBO J 12 199320052015
    Crossref | PubMed | ISI | Google Scholar
  • 390 Migliaccio E, Giorgio M, Mele S, Pelicci G, Reboldi P, Pandolfi PP, Lanfrancone L, Pelicci PG The p66shc adaptor protein controls oxidative stress response and life span in mammals. Nature 402 1999309313
    Crossref | PubMed | ISI | Google Scholar
  • 391 Mihm S, Ennen J, Pessara U, Kurth R, Dröge W Inhibition of HIV-1 replication and NFκB activity by cysteine and cysteine derivatives. AIDS 5 1991497503
    Crossref | PubMed | ISI | Google Scholar
  • 392 Milhavet O, McMahon HEM, Rachidi W, Nishida N, Katamine S, Mangé A, Arlotto M, Casanova D, Riondel J, Favier A, Lehmann S Prion infection impairs the cellular response to oxidative stress. Proc Natl Acad Sci USA 97 20001393713942
    Crossref | PubMed | ISI | Google Scholar
  • 393 Miquel J, Ferrandiz ML, de Juan E, Sevila I, Martinez M N-acetylcysteine protects against age-related decline of oxidative phosphorylation in liver mitochondria. Eur J Pharmacol 292 1995333335
    PubMed | ISI | Google Scholar
  • 394 Mittal CK, Murad F Activation of guanylate cyclase by superoxide dismutase and hydroxyl radical: a physiological regulator of guanosine 3′,5′-monophosphate formation. Proc Natl Acad Sci USA 74 197743604364
    Crossref | PubMed | ISI | Google Scholar
  • 395 Mohammadi M, Schlessinger J, Hubbard SR Structure of the FGF receptor tyrosine kinase domain reveals a novel autoinhibitory mechanism. Cell 86 1996577587
    Crossref | PubMed | ISI | Google Scholar
  • 396 Mohazzab KM, Kaminski PM, Wolin MS NADH oxidoreductase is a major source of superoxide anion in bovine coronary artery endothelium. Am J Physiol Heart Circ Physiol 266 1994H2568H2572
    Link | Google Scholar
  • 397 Monteiro HP, Stern A Redox modulation of tyrosine phosphorylation-dependent signal transduction pathways. Free Radical Biol Med 21 1996323333
    Crossref | PubMed | ISI | Google Scholar
  • 398 Montine KS, Kim PJ, Olson SJ, Markesbery WR, Montine TJ 4-Hydroxy-2-nonenal pyrrole adducts in human neurodegenerative disease. J Neuropathol Exp Neurol 56 1997866871
    Crossref | PubMed | ISI | Google Scholar
  • 399 Morgan BA, Banks GR, Toone WM, Raitt D, Kuge S, Johnston LH The Skn7 response regulator controls gene expression in the oxidative stress response of the budding yeast Saccharomyces cerevisiae. EMBO J 16 199710351044
    Crossref | PubMed | ISI | Google Scholar
  • 400 Morgenstern DE, Gifford MAC, Li LL, Doerschuk CM, Dinauer MC Absence of respiratory burst in X-linked chronic granulomatous disease mice leads to abnormalities in both host defense and inflammatory response to Aspergillus fumigatus. J Exp Med 185 1997207218
    Crossref | PubMed | ISI | Google Scholar
  • 401 Morris BJ Stumulation of immediate early gene expression in striatal neurons by nitric oxide. J Biol Chem 270 19952474024744
    PubMed | ISI | Google Scholar
  • 402 Muehlematter D, Ochi T, Cerutti P Effects of tert-butyl hydroperoxide on promotable and non-promotable JB6 mouse epidermal cells. Chem-Biol Interact 71 1989339352
    Crossref | PubMed | ISI | Google Scholar
  • 403 Mukherjee SP, Attaway EJ, Mukerjee C Insulin-like stimulation by hydrogen peroxide production in adipocyte by insulin receptor antibodies. Biochem Int 4 1982305
    Google Scholar
  • 404 Multhaup G, Ruppert T, Schlicksupp A, Hesse L, Beher D, Masters CL, Beyreuther K Reactive oxygen species and Alzheimer's disease. Biochem Pharmacol 54 1997533539
    Crossref | PubMed | ISI | Google Scholar
  • 405 Munro JM Endothelial-leukocyte adhesive interactions in inflammatory diseases. Eur Heart J 14 Suppl K 19937277
    PubMed | ISI | Google Scholar
  • 406 Munroe DG, Wang EY, MacIntyre JP, Tam SSC, Lee DHS, Taylor GR, Zhou L, Plante RK, Kazmi SMI, Baeuerle PA, Lau CY Novel intracellular signalling function of prostaglandin H synthase-1 in NF-κB activation. J Inflamm 45 1995260268
    PubMed | Google Scholar
  • 407 Nakamura H, Matsuda M, Furuke K, Kitaoka Y, Iwata S, Toda K, Inamoto T, Yamaoka Y, Ozawa K, Yodoi J Adult T cell leukemia-derived factor/human thioredoxin protects endothelial F-2 cell injury caused by activated neutrophils or hydrogen peroxide. Immunol Lett 42 19947580
    Crossref | PubMed | ISI | Google Scholar
  • 408 Nakamura H, Nakamura K, Yodoi J Redox regulation of cellular activation. Annu Rev Immunol 15 1997351369
    Crossref | PubMed | ISI | Google Scholar
  • 409 Nakamura K, Hori T, Sato N, Sugie K, Kawakami T, Yodoi J Redox regulation of a Src family protein tyrosine kinase p56lck in T cells. Oncogene 8 199331333139
    PubMed | ISI | Google Scholar
  • 410 Nakamura Y, Gindhart T, Winterstein D, Tomita I, Seed J, Colburn N Early superoxide dismutase-sensitive event promotes neoplastic transformation in mouse epidermal JB6 cells. Carcinogenesis 9 1988203207
    Crossref | PubMed | ISI | Google Scholar
  • 411 Nakashima I, Pu M, Nishizaki A, Rosila I, Ma L, Katano Y, Ohkusu K, Rahman SMJ, Isobe K, Hamaguchi M, Saga K Redox mechanism as alternaive to ligand binding for receptor activation delivering disregulated cellular signals. J Immunol 152 199410641071
    PubMed | ISI | Google Scholar
  • 412 Nathan CF, Root RK Hydrogen peroxide release from mouse peritoneal macrophages: dependence on sequential activation and triggering. J Exp Med 146 197716481662
    Crossref | PubMed | ISI | Google Scholar
  • 413 Neumcke I, Schneider B, Fandrey J, Pagel H Effects of pro- and antioxidative compounds on renal production of erythropoietin. Endocrinology 140 1999641645
    Crossref | PubMed | ISI | Google Scholar
  • 414 Newman E, Heslin MJ, Wolf RF, Pisters PW, Brennan MF The effect of insulin on glucose and protein metabolism in the forearm of cancer patients. Surg Oncol 1 1992257267
    Crossref | PubMed | Google Scholar
  • 415 Nishikawa T, Edelstein D, Du XL, Yamagishi S, Matsumura T, Kaneda Y, Yorek MA, Beebe D, Oates PJ, Hammes HP, Giardino I, Brownlee M Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature 404 2000787790
    Crossref | PubMed | ISI | Google Scholar
  • 416 Nishizuka Y Intracellular signaling by hydrolysis of phospholipids and activation of protein kinase C. Science 258 1992607614
    Crossref | PubMed | ISI | Google Scholar
  • 417 Nohl H, Gille L, Schönheit K, Liu Y Conditions allowing redox-cycling ubisemiquinone in mitochondria to establish a direct redox couple with molecular oxygen. Free Radical Biol Med 20 1996207213
    Crossref | PubMed | ISI | Google Scholar
  • 418 Nohl H, Hegner D Do mitochondria produce oxygen free radicals in vivo? Eur J Biochem 82 1978563567
    Crossref | PubMed | Google Scholar
  • 419 Nose K, Shibanuma M, Kikuchi K, Kageyama H, Sakiyama S, Kuroki T Transcriptional activation of early-response genes by hydrogen peroxide in a mouse osteoblastic cell line. Eur J Biochem 201 199199106
    Crossref | PubMed | Google Scholar
  • 420 Oberley LW Free radicals and diabetes. Free Radical Biol Med 5 1988113124
    Crossref | PubMed | ISI | Google Scholar
  • 421 O'Brien RM, Granner DK Regulation of gene expression by insulin. Physiol Rev 76 199611091161
    Link | ISI | Google Scholar
  • 422 Odom AL, Hatwig CA, Stanley JS, Benson AM Biochemical determinants of Adriamycin toxicity in mouse liver, heart and intestine. Biochem Pharmacol 43 1992831836
    Crossref | PubMed | ISI | Google Scholar
  • 423 O'Donnell VB, Spycher S, Azzi A Involvement of oxidants and oxidant-generating enzyme(s) in tumor necrosis factor-α-mediated apoptosis: role for lipoxygenase pathway but not mitochondrial respiratory chain. Biochem J 310 1995133141
    Crossref | PubMed | ISI | Google Scholar
  • 424 Ohara Y, Peterson TE, Harrison DG Hypercholesterolemia increases endothelial superoxide anion production. J Clin Invest 91 199325462551
    Crossref | PubMed | ISI | Google Scholar
  • 425 Ohga E, Nagase T, Tomita T, Teramoto S, Matsuse T, Katayama H, Ouchi Y Increased levels of circulating ICAM-1, VCAM-1, and L-selectin in obstructive sleep apnea syndrome. J Appl Physiol 87 19991014
    Link | ISI | Google Scholar
  • 426 Ohki K, Yoshida K, Hagiwara M, Harada T, Takamura M, Ohashi T, Matsuda H, Imaki J Nitric oxide induces c-fos gene expression via cyclic AMP response element binding protein (CREB) phosphorylation in rat retinal pigment epithelium. Brain Res 696 1995140144
    Crossref | PubMed | ISI | Google Scholar
  • 427 Okabe E, Kato Y, Sasaki H, Saito G, Hess ML, Ito H Calmodulin participation in oxygen radical-induced sarcoplasmic reticulum calcium uptake reduction. Arch Biochem Biophys 255 1987464468
    Crossref | PubMed | ISI | Google Scholar
  • 428 Okabe E, Sugihara M, Tanaka K, Sasaki H, Ito H Calmodulin and free oxygen radicals interaction with steady-state calcium accumulation and passive calcium permeability of cardiac sarcoplasmic reticulum. J Pharmacol Exp Ther 250 1989286292
    PubMed | ISI | Google Scholar
  • 429 Okabe E, Kuse K, Sekishita T, Suyama N, Tanaka K, Ito H The effect of ryanodine on oxygen free radical-induced dysfunction of cardiac sarcoplasmic reticulum. J Pharmacol Exp Ther 256 1991868875
    PubMed | ISI | Google Scholar
  • 430 Ono Y, Fujii T, Ogita K, Kikkawa U, Igarishi K, Nishizuka Y Protein kinase C zeta subspecies from rat brain: its structure, expression, and properties. Proc Natl Acad Sci USA 86 198930993103
    Crossref | PubMed | ISI | Google Scholar
  • 431 Orr WC, Sohal RS Extension of lifespan by overexpression of superoxide dismutase and catalase in Drosophila melanogaster. Science 263 199411281130
    Crossref | PubMed | ISI | Google Scholar
  • 432 Otani H, Engelman RM, Rousou JA, Breyer RH, Das DK Enhanced prostaglandin synthesis due to phospholipase breakdown in ischemic reperfused myocardium. Control of its production by a phospholipase inhibitor or free radical scavengers. J Mol Cell Cardiol 18 1986953961
    Crossref | PubMed | ISI | Google Scholar
  • 433 Otani H, Engelman RM, Rousou JA, Breyer RH, Lemeshow S, Das DK Cardiac performance during reperfusion improved by pretreatment with oxygen-free radical scavengers. J Thorac Cardiovasc Surg 91 1986290295
    PubMed | ISI | Google Scholar
  • 434 Ozaki M, Deshpande SS, Angkeow P, Bellan J, Lowenstein CJ, Dinauer MC, Goldschmidt-Clermont PJ, Irani K Inhibition of the Rac1 GTPase protects against nonlethal ischemia/reperfusion-induced necrosis and apoptosis in vivo. FASEB J 14 2000418429
    Crossref | PubMed | ISI | Google Scholar
  • 435 Packer L, Fuehr K Low oxygen concentration extends the lifespan of cultured human diploid cells. Nature 267 1977423425
    Crossref | PubMed | ISI | Google Scholar
  • 436 Pahlavani MA, Harris MD Effect of in vitro generation of oxygen free radicals on T cell function in young and old rats. Free Radical Biol Med 25 1998903913
    Crossref | PubMed | ISI | Google Scholar
  • 437 Palmer RMJ, Rees DD, Ashton DS, Moncada S l-Arginine is the physiological precursor for the formation of nitric oxide in endothelium dependent relaxation. Biochem Biophys Res Commun 153 198812511256
    Crossref | PubMed | ISI | Google Scholar
  • 438 Pantel K, Schlimok G, Angstwurm M, Passlick B, Izbicki JR, Johnson JP, Riethmuller G Early metastasis of human solid tumors: expression of cell adhesion molecules. Ciba Found Symp 189 1995157170
    PubMed | Google Scholar
  • 439 Parkes TL, Elia AJ, Dickinson D, Hilliker AJ, Boulianne GL, John P Extension of Drosophila lifespan by overexpression of human SOD1 in motorneurons. Nature Genet 19 1998171174
    Crossref | PubMed | ISI | Google Scholar
  • 440 Pasinelli P, Houseweart MK, Brown RH, Cleveland DW Caspase-1 and -3 are sequentially activated in motor neuron death in Cu,Zn superoxide dismutase-mediated familial amyotrophic lateral sclerosis. Proc Natl Acad Sci USA 97 20001390113906
    Crossref | PubMed | ISI | Google Scholar
  • 441 Pauly PC, Harris DA Copper stimulates endocytosis of the prion protein. J Biol Chem 273 19983310733110
    Crossref | PubMed | ISI | Google Scholar
  • 442 Pearson M, Carbone R, Sebastiani C, Cioce M, Fagioli M, Saito S, Higashimoto Y, Appella E, Minucci S, Pandolfi PP, Pelicci PG PML regulates p53 acetylation and premature senescence induced by oncogenic Ras. Nature 406 2000207210
    Crossref | PubMed | ISI | Google Scholar
  • 443 Peker Y, Hedner J, Kraiczi H, Löth S Respiratory disturbance index. Am J Respir Crit Care Med 162 20008186
    Crossref | PubMed | ISI | Google Scholar
  • 444 Peppard PE, Young T, Palta M, Skatrud J Prospective study of the association between sleep-disordered breathing and hypertension. N Engl J Med 342 200013781383
    Crossref | PubMed | ISI | Google Scholar
  • 445 Pfeilschifter J, Huwiler A Nitric oxide stimulates stress-activated protein kinases in glomerular endothelial and mesangial cells. FEBS Lett 396 19966770
    Crossref | PubMed | ISI | Google Scholar
  • 446 Piette J, Piret B, Bonizzi G, Schoonbroodt S, Merville MP, Legrand-Poels S, Bours B Multiple redox regulation in NF-κB transcription factor activation. Biol Chem 378 199712371245
    PubMed | ISI | Google Scholar
  • 447 Pisters PWT, Cersosimo E, Rogatko A, Brennan MF Insulin action on glucose and branched-chain amino acid metabolism in cancer cachexia: differential effects of insulin. Surgery 111 1992301310
    PubMed | ISI | Google Scholar
  • 448 Pollock JD, Williams DA, Gifford MA, Li LL, Du X, Fisherman J, Orkin SH, Doerschuk CM, Dinauer MC Mouse model of X-linked chronic granulomatous diseases, an inherited defect in phagocyte superoxide production. Nat Genet 9 1995202209
    Crossref | PubMed | ISI | Google Scholar
  • 449 Prabhakar NR Oxygen sensing by the carotid body chemoreceptors. J Appl Physiol 88 200022872295
    Link | ISI | Google Scholar
  • 450 Praticò D, Lee VMY, Trojanowski JQ, Rokach J, Fitzgerald GA Increased F2-isoprostanes in Alzheimer's disease: evidence for enhanced lipid peroxidation in vivo. FASEB J 12 199817771783
    Crossref | PubMed | ISI | Google Scholar
  • 451 Prusiner SB, Scott MR, DeArmond SJ, Cohen FE Prion protein biology. Cell 93 1998337348
    Crossref | PubMed | ISI | Google Scholar
  • 452 Pu M, Akhand AA, Kato M, Hamaguchi M, Koike T, Iwata H, Sabe H, Suzuki H, Nakashima I Evidence of a novel redox-linked activation mechanism for the Src kinase which is independent of tyrosine 527-mediated regulation. Oncogene 13 199626152622
    PubMed | ISI | Google Scholar
  • 453 Qin J, Clore GM, Kennedy WM, Huth JR, Gronenborn AM Solution structure of human thioredoxin in a mixed disulfide intermediate complex with its target peptide from the transcription factor NF-κB. Structure 3 1995289297
    Crossref | PubMed | ISI | Google Scholar
  • 454 Qu CK, Yu WM, Azzarelli B, Feng GS Genetic evidence that shp-2 tyrosine phosphatase is a signal enhancer of the epidermal growth factor receptor in mammals. Proc Natl Acad Sci USA 96 199985288533
    Crossref | PubMed | ISI | Google Scholar
  • 455 Quest A, Bardes ES, Bell RM A phorbol ester binding domain of protein kinase C gamma. Deletion analysis of the cys2 domain defines a minimal 43 amino acid peptide. J Biol Chem 269 199429612970
    PubMed | ISI | Google Scholar
  • 456 Radomski MW, Palmer RMJ, Moncada S The anti-aggregating properties of vascular endothelium: interactions between prostacyclin and nitric oxide. Br J Pharmacol 92 1987639646
    Crossref | PubMed | ISI | Google Scholar
  • 457 Rando RR, Kishi Y Structural basis of protein kinase C activation. Biochemistry 31 199222112218
    Crossref | PubMed | ISI | Google Scholar
  • 458 Rao GN Hydrogen peroxide induces complex formation of SHC-Grb2-SOS with receptor tyrosine and activates Ras and extracellular signal-regulated protein kinases group of mitogen-activated protein kinases. Oncogene 13 1996713719
    PubMed | ISI | Google Scholar
  • 459 Reid MB, Haack KE, Franchek KM, Valberg PA, Kobzik L, West MS Reactive oxygen in skeletal muscle. I. Intracellular oxidant kinetics and fatigue in vitro. J Appl Physiol 73 199217971804
    Link | ISI | Google Scholar
  • 460 Reid MB, Stokic DS, Koch SM, Khawli FA, Leis AA N-acetylcysteine inhibits muscle fatigue in humans. J Clin Invest 94 199424682474
    Crossref | PubMed | ISI | Google Scholar
  • 461 Reid VC, Mitchinson MJ, Skepper JN Cytotoxicity of oxidised low-density lipoprotein to mouse peritoneal macrophages: an ultrastructural study. J Pathol 171 1993321328
    Crossref | PubMed | ISI | Google Scholar
  • 462 Res PCM, Schaar CG, Breedveld FC, van Eden W, van Embden JDA, Cohen IR, de Vries RRP Synovial fluid T cell reactivity against the 65 kD heat-shock protein of mycobacteria in early onset of chronic arthritis. Lancet 2 1988478480
    Crossref | PubMed | ISI | Google Scholar
  • 463 Reznick AZ, Packer L, Sen CK Strategies to assess oxidative stress. Oxidative Stress in Skeletal Muscle , Reznick AZ 1998Birkhäuser VerlagBasel, p.43–58.
    Google Scholar
  • 464 Riabowol K, Schiff J, Gilman MZ Transcription factor AP-1 activity is required for initiation of DNA synthesis and is lost during cellular aging. Proc Natl Acad Sci USA 89 1992157161
    Crossref | PubMed | ISI | Google Scholar
  • 465 Richman PG, Meister A Regulation of γ-glutamyl-cysteine synthetase by nonallosteric feedback inhibition by glutathione. J Biol Chem 250 197514221426
    PubMed | ISI | Google Scholar
  • 466 Roberts AW, Kim C, Zhen L, Lowe JB, Kapur R, Petryniak B, Spaetti A, Pollok JD, Borneo JB, Bradford GB, Atkinson SJ, Dinauer MC, Williams DA Deficiency of the hematopoietic cell-specific Rho family GTPase Rac2 is characterized by abnormalities in neutrophil function and host defense. Immunity 10 1999183196
    Crossref | PubMed | ISI | Google Scholar
  • 467 Rodemann HP, Bayreuther K, Francz PI, Dittmann K, Albiez M Selective enrichment and biochemical characterization of seven human skin fibroblasts cell types in vitro. Exp Cell Res 180 19898493
    Crossref | PubMed | ISI | Google Scholar
  • 468 Roederer M, Staal FJT, Osada H, Herzenberg LA, Herzenberg LA CD4 and CD8 T cells with high intracellular glutathione levels are selectively lost as the HIV infection progresses. Int Immunol 3 1991933937
    Crossref | PubMed | ISI | Google Scholar
  • 469 Rohdenburg GL, Bernhard A, Krehbiel O Sugar tolerance in cancer. JAMA 72 191915281534
    Crossref | Google Scholar
  • 470 Rosen DR, Siddique T, Patterson D, Figlewicz DA, Sapp P, Hentati A, Donaldson D, Goto J, O'Regan JP, Deng H-X, Rahmani Z, Krizus A, McKenna-Yasek D, Cayabyab A, Gaston SM, Berger R, Tanzi RE, Halperin JJ, Herzfeldt B, Van den Bergh R, Hung W-Y, Bird T, Deng G, Mulder DW, Smyth C, Laing NG, Soriano E, Pericak-Vance MA, Haines J, Rouleau GA, Gusella JS, Horvitz HR, Brown RH Mutations in Cu/Zn SOD gene are associated with familial amyotrophic lateral sclerosis. Nature 362 19935962
    Crossref | PubMed | ISI | Google Scholar
  • 471 Rosen OM, Herrera R, Olowe Y, Petruzzelli LM, Cobb MH Phosphorylation activates the insulin receptor tyrosine protein kinase. Proc Natl Acad Sci USA 80 198332373240
    Crossref | PubMed | ISI | Google Scholar
  • 472 Roth S, Dröge W Regulation of T cell activation and T cell growth factor (TCGF) production by hydrogen peroxide. Cell Immunol 108 1987417424
    Crossref | PubMed | ISI | Google Scholar
  • 473 Roth S, Dröge W Regulation of interleukin-2 production, interleukin 2 mRNA expression and intracellular glutathione levels in ex vivo derived T lymphocytes by lactate. Eur J Immunol 21 199119331937
    Crossref | PubMed | ISI | Google Scholar
  • 474 Roth S, Dröge W Glutathione reverses the inhibition of T cell responses by superoptimal numbers of "nonprofessional" antigen presenting cells. Cell Immunol 155 1994183194
    Crossref | PubMed | ISI | Google Scholar
  • 475 Roth S, Gmünder H, Dröge W Regulation of intracellular glutathione levels and lymphocyte functions by lactate. Cell Immunol 136 199195104
    Crossref | PubMed | ISI | Google Scholar
  • 476 Roveri A, Coassin M, Maiorino M, Zamburlini A, van Amsterdam FT, Ratti E, Ursini F Effect of hydrogen peroxide on calcium homeostasis in smooth muscle cells. Arch Biochem Biophys 297 1992265270
    Crossref | PubMed | ISI | Google Scholar
  • 477 Roy S, Sen CK, Gozin A, Andrieu V, Pasquier C Redox regulation of cell adhesion processes. Antioxidants and Redox Regulation of Genes , Sen K, Sies H, Baeuerle P 2000265295AcademicSan Diego, CA
    Google Scholar
  • 478 Roy S, Sen CK, Packer L Determination of cell-cell adhesion in response to oxidants and antioxidants. Methods in Enzymology: Oxidants and Antioxidants. 1999395401AcademicSan Diego, CA
    Google Scholar
  • 479 Saavedra JE, Billiar TR, Williams DL, Kim YM, Watkins SC, Keefer LK Targeting nitric oxide (NO) delivery in vivo. Design of a liver-selective NO donor prodrug that blocks tumor necrosis factor-α-induced apoptosis and toxicity in the liver. J Med Chem 40 199719471954
    Crossref | PubMed | ISI | Google Scholar
  • 480 Sachi Y, Hirota K, Masutani H, Toda K, Okamoto T, Takigawa M, Yodoi J Induction of ADF/TRX by oxidative stress in keratinocytes and lymphoid cells. Immunol Lett 44 1995189193
    Crossref | PubMed | ISI | Google Scholar
  • 481 Saitoh M, Nishitoh H, Fujii M, Takeda K, Tobiume K, Sawada Y, Kawabata M, Miyazono K, Ichijo H Mammalian thioredoxin is a direct inhibitor of apoptosis signal-regulating kinase (ASK)1. EMBO J 17 199825962606
    Crossref | PubMed | ISI | Google Scholar
  • 482 Sakurai T, Tsuchiya S Superoxide production from nonenzymatically glycated protein. FEBS Lett 236 1988406410
    Crossref | PubMed | ISI | Google Scholar
  • 483 Salim AS The permissive role of oxygen-derived free radicals in the development of colonic cancer in the rat. A new theory for carcinogenesis. Int J Cancer 53 199310311035
    Crossref | PubMed | ISI | Google Scholar
  • 484 Salvemini D, Wang ZQ, Zweier JL, Samouilov A, MacArthur H, Misko TP, Currie MG, Cuzzocrea S, Sikorski JA, Riley DP A nonpeptidyl mimic of superoxide dismutase with therapeutic activity in rats. Science 286 1999304306
    Crossref | PubMed | ISI | Google Scholar
  • 485 Sastre J, Asensi M, Gascó E, Pallardó FV, Ferrero JA, Furukawa T, Vina J Exhaustive physical exercise causes oxidation of glutathione status in blood: prevention by antioxidant administration. Am J Physiol Regulatory Integrative Comp Physiol 263 1992R992R995
    Link | ISI | Google Scholar
  • 486 Sato H, Tamba M, Ishii T, Bannai S Cloning and expression of a plasma membrane cystine/glutamate exchange transporter composed of two distinct proteins. J Biol Chem 274 19991145511458
    Crossref | PubMed | ISI | Google Scholar
  • 487 Sayre LM, Zelasko DA, Harris PLR, Perry G, Salomon RG, Smith MA 4-Hydroxynonenal-derived advanced lipid peroxidation end products are increased in Alzheimer's disease. J Neurochem 68 199720922097
    Crossref | PubMed | ISI | Google Scholar
  • 488 Schaller MD, Borgman CA, Cobb BS, Vines RR, Reynolds AB, Parsons JT pp125fak a structurally distinctive protein-tyrosine kinase associated with focal adhesions. Proc Natl Acad Sci USA 89 199251925196
    Crossref | PubMed | ISI | Google Scholar
  • 489 Schell MA Molecular biology of the LysR family of transcriptional regulators. Annu Rev Microbiol 47 1993597626
    Crossref | PubMed | ISI | Google Scholar
  • 490 Schenk H, Klein M, Erdbrügger W, Dröge W, Schulze-Osthoff K Distinct effects of thioredoxin and antioxidants on the activation of transcription factors NF-κB and AP-1. Proc Natl Acad Sci USA 91 199416721676
    Crossref | PubMed | ISI | Google Scholar
  • 491 Schieven GL, Kirihara JM, Burg DL, Geahlen RL, Ledbetter JA p72syk tyrosine kinase is activated by oxidizing conditions that induce lymphocyte tyrosine phosphorylation and Ca2+ signals. J Biol Chem 268 19931668816692
    PubMed | ISI | Google Scholar
  • 492 Schieven GL, Kirihara JM, Myers DE, Ledbetter JA, Uckun FM Reactive oxygen intermediates activate NF-kappa B in a tyrosine kinase-dependent mechanism and in combination with vanadate activate the p56lck and p59fyn tyrosine kinases in human lymphocytes. Blood 82 199312121220
    PubMed | ISI | Google Scholar
  • 493 Schieven GL, Mittler RS, Nadler SG, Kirihara JM, Bolen JB, Kanner SB, Ledbetter JA ZAP-70 tyrosine kinase, CD45, and T cell receptor involvement in UV- and H2O2-induced T cell signal transduction. J Biol Chem 269 19942071820726
    PubMed | ISI | Google Scholar
  • 494 Schmid E, El Benna J, Galter D, Klein G, Dröge W Redox priming of the insulin receptor β-chain associated with altered tyrosine kinase activity and insulin responsiveness in the absence of tyrosine autophosphorylation. FASEB J 12 1998863870
    Crossref | PubMed | ISI | Google Scholar
  • 495 Schmid E, Hotz-Wagenblatt A, Dröge W Inhibition of the insulin receptor kinase phosphorylation by nitric oxide: functional and structural aspects. Antioxidants Redox Signal 1 19994553
    Crossref | PubMed | ISI | Google Scholar
  • 496 Schmid E, Hotz-Wagenblatt A, Hack V, Dröge W Phosphorylation of the insulin receptor kinase by phosphocreatine in combination with hydrogen peroxide. The structural basis of redox priming. FASEB J 13 199914911500
    Crossref | PubMed | ISI | Google Scholar
  • 497 Schmidt HHHW, Lohmann SM, Walter U The nitric oxide and cGMP signal transduction system: regulation and mechanism of action. Biochim Biophys Acta 1178 1993153175
    Crossref | PubMed | ISI | Google Scholar
  • 498 Schnell N, Krems B, Entian KD The PAR1 (YAP1/SNQ3) gene of Saccharomyces cerevisiae, a c-jun homologue, is involved in oxygen metabolism. Curr Genet 21 1992269273
    Crossref | PubMed | ISI | Google Scholar
  • 499 Schoonbroodt S, Ferreira V, Best-Belpomme M, Boelaert JR, Legrand-Poels S, Korner M, Piette J Crucial role of the amino-terminal tyrosine residue 42 and the carboxyl-terminal PEST domain of IκBα in NF-κB activation by an oxidative stress. J Immunol 164 200042924300
    Crossref | PubMed | ISI | Google Scholar
  • 500 Schoonbroodt S, Legrand-Poels S, Best-Belpomme M, Piette J Activation of the NF-κB transcription factor in a T-lymphocytic cell line by hypochlorous acid. Biochem J 321 1997777785
    Crossref | PubMed | ISI | Google Scholar
  • 501 Schreck R, Baeuerle PA Reactive oxygen intermediates as apparently widely used messengers in the activation of NF-κB transcription factor and HIV-1. Trends Cell Biol 1 19913942
    Crossref | PubMed | Google Scholar
  • 502 Schreck R, Rieber P, Baeuerle PA Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-κB transcription factor and HIV-1. EMBO J 10 199122472258
    Crossref | PubMed | ISI | Google Scholar
  • 503 Schulz R, Mahmoudi S, Hattar K, Sibelius U, Olschewski H, Mayer K, Seeger W, Grimminger F Enhanced release of superoxide from polymorphonuclear neutrophils in obstructive apnea. Am J Respir Crit Care Med 162 2000566570
    Crossref | PubMed | ISI | Google Scholar
  • 504 Schulze-Osthoff K, Bakker AC, Vanhaesebroeck B, Beyaert R, Jacob WA, Fiers W Cytotoxic activity of tumor necrosis factor is mediated by early damage of mitochondrial functions. Evidence for the involvement of mitochondrial radical genration. J Biol Chem 267 199253175323
    PubMed | ISI | Google Scholar
  • 505 Schulze-Osthoff K, Beyaert R, Vandevoorde V, Haegeman G, Fiers W Depletion of the mitochondrial transport abrogates the cytotoxic and gene-inductive effects of TNF. EMBO J 12 199330953104
    Crossref | PubMed | ISI | Google Scholar
  • 506 Schulze-Osthoff K, Schenk H, Dröge W Effects of thioredoxin on activation of transcription factor NF-κB. Methods Enzymol 252 1995253264
    Crossref | PubMed | ISI | Google Scholar
  • 507 Schwartz RH Costimulation of T lymphocytes: the role of CD28, CTLA-4, and B7/BB1 in interleukin-2 production and immunotherapy. Cell 71 199210651068
    Crossref | PubMed | ISI | Google Scholar
  • 508 Segal J Aging: a nonregulated process. Med Hypotheses 26 1988197207
    Crossref | ISI | Google Scholar
  • 509 Sellak H, Franzini E, Hakim J, Pasquier C Reactive oxygen species rapidly increase endothelial ICAM-1 ability to bind neutrophils without detectable upregulation. Blood 83 199426692677
    Crossref | PubMed | ISI | Google Scholar
  • 510 Semenza GL HIF-1: mediator of physiological and pathophysiological responses to hypoxia. J Appl Physiol 88 200014741480
    Link | ISI | Google Scholar
  • 511 Sen CK, Atalay M, Hänninen O Exercise-induced oxidative stress: glutathione supplementation and deficiency. J Appl Physiol 77 199421772187
    Link | ISI | Google Scholar
  • 512 Sen CK, Packer L Antioxidant and redox regulation of gene transcription. FASEB J 10 1996709720
    Crossref | PubMed | ISI | Google Scholar
  • 513 Sen CK, Rankinen T, Väisänen S, Rauramaa RR Oxidative stress after human exercise: effect of N-acetyl-cysteine supplementation. J Appl Physiol 76 199425702577
    Link | ISI | Google Scholar
  • 514 Serrano CV, Mikhail EA, Wang P, Noble B, Kuppusamy P, Zweier JL Superoxide and hydrogen peroxide induce CD18-mediated adhesion in the postischemic heart. Biochim Biophys Acta 1316 1996191202
    Crossref | PubMed | ISI | Google Scholar
  • 515 Shalaby MR, Aggarwal BB, Rinderknecht E, Svedersky LP, Finkle BS, Palladino MA Activation of human polymorphonuclear neutrophil functions by interferon-γ and tumor necrosis factor. J Immunol 135 198520692073
    PubMed | ISI | Google Scholar
  • 516 Shapiro VS, Truitt KE, Imboden JB, Weiss A CD28 mediates transcriptional upregulation of the interleukin-2 (IL-2). Promoter through a composite element containing the CD28RE and NF-IL-2B AP-1 sites. Mol Cell Biol 17 199740514058
    Crossref | PubMed | ISI | Google Scholar
  • 517 Shaw JH, Klein F, Wolfe RR Assessment of alanine, urea and glucose interrelationships in normal subjects and in patients with sepsis using stable isotopic tracers. Surgery 97 1985557567
    PubMed | ISI | Google Scholar
  • 518 Shaw JH, Wolfe RR Glucose and urea kinetics in patients with early and advanced gastrointestinal cancer: the response to glucose infusion and TPN. Surgery 101 1987181186
    PubMed | ISI | Google Scholar
  • 519 Shepard JW Hypertension, cardiac arrhythmias, myocardial infarction and stroke in relation to obstructive sleep apnea. Clin Chest Med 13 1992437458
    PubMed | ISI | Google Scholar
  • 520 Shibanuma M, Kuroki T, Nose K Induction of DNA replication and expression of proto-oncogene c-myc and c-fos in quiescent Balb/3T3 cells by xanthine/xanthine oxidase. Oncogene 3 19881721
    ISI | Google Scholar
  • 521 Shiloh MU, MacMicking JD, Nicholson S, Brause JE, Potter S, Marino M, Fang F, Dinauer M, Nathan C Phenotype of mice and macrophages deficient in both phagocyte oxidase and inducible nitric oxide synthase. Immunity 10 19992938
    Crossref | PubMed | ISI | Google Scholar
  • 522 Shull S, Heintz NH, Periasamy M, Manohar M, Janssen YM, Marsh JP, Mossman BT Differential regulation of antioxidant enzymes in response to oxidants. J Biol Chem 266 19912439824403
    PubMed | ISI | Google Scholar
  • 523 Sicheri F, Moarefi I, Kuriyan J Crystal structure of the Src family tyrosine kinase Hck. Nature 385 1997602609
    Crossref | PubMed | ISI | Google Scholar
  • 524 Sido B, Braunstein J, Breitkreutz R, Herfarth C, Meuer SC Thiol-mediated redox regulation of intestinal lamina propria T lymphocytes. J Exp Med 192 2000907912
    Crossref | PubMed | ISI | Google Scholar
  • 525 Siebenlist U, Franzoso G, Brown K Structure, regulation and function of NFκB. Annu Rev Cell Biol 10 1994405455
    Crossref | PubMed | Google Scholar
  • 526 Sies H Biochemie des oxidativen Stress. Angew Chem 98 198610611075
    Crossref | Google Scholar
  • 527 Sies H Strategies of antioxidant defense. Eur J Biochem 215 1993213219
    Crossref | PubMed | Google Scholar
  • 528 Siman CM, Eriksson UJ Vitamin E decreases the occurrence of malformations in the offspring of diabetic rats. Diabetes 46 199710541061
    Crossref | PubMed | ISI | Google Scholar
  • 529 Sinet PM, Coutrier J, Dutillaux B Trisomie 21 et superoxyde dismutase-1 (IPO-A): tentative de localisation sur la sous-bande 21q22.1. Exp Cell Res 97 19764755
    Crossref | PubMed | ISI | Google Scholar
  • 530 Sitte N, Huber M, Grune T, Ladhoff A, Doecke WD, von Zglinicki T, Davies KJA Proteasome inhibition by lipofuscin/ceroid during postmitotic aging of fibroblasts. FASEB J 14 200014901498
    Crossref | PubMed | ISI | Google Scholar
  • 531 Sitte N, Merker K, von Zglinicki T, Davies KJA, Grune T Protein oxidation and degradation during cellular senescence of human BJ-fibroblasts. Part II. Aging of non-dividing cells. FASEB J 14 200025032510
    Crossref | PubMed | ISI | Google Scholar
  • 532 Sitte N, Merker K, von Zglinicki T, Grune T, Davies KJA Protein oxidation and degradation during cellular senescence of human BJ-fibroblasts. Part I. Effects of proliferative senescence. FASEB J 14 200024952502
    Crossref | PubMed | ISI | Google Scholar
  • 533 Sitte N, Saretzki G, von Zglinicki T Accelerated telomere shortening in fibroblasts after extended periods of confluency. Free Radical Biol Med 24 1997885893
    Crossref | ISI | Google Scholar
  • 534 Sjödin B, Hellsten Westing Y, Apple FS Biochemical mechanisms for oxygen free radical formation during exercise. Sports Med 10 1990236254
    Crossref | PubMed | ISI | Google Scholar
  • 535 Slater AF, Stefan C, Novel I, van den Dobbelsteen DJ, Orrenius S Signalling mechanisms and oxidative stress in apoptosis. Toxicol Lett 82–83 1995149153
    Crossref | PubMed | ISI | Google Scholar
  • 536 Smith J, Ladi E, Mayer-Pröschel M, Noble M Redox state is a central modulator of the balance between self-renewal and differentiation in a dividing glial precursor cell. Proc Natl Acad Sci USA 97 20001003210037
    Crossref | PubMed | ISI | Google Scholar
  • 537 Sohal RS, Dubey A Mitochondrial oxidative damage, hydrogen peroxide release, and aging. Free Radical Biol Med 16 1994621626
    Crossref | PubMed | ISI | Google Scholar
  • 538 Sohal RS, Ku HH, Agarwal S, Forster MJ, Lal H Oxidative damage, mitochondrial oxidant generation and antioxidant defenses during aging and in response to food restriction in the mouse. Mech Ageing Dev 74 1994121133
    Crossref | PubMed | ISI | Google Scholar
  • 539 Sohal RS, Weindruch R Oxidative stress, caloric restriction, and aging. Science 273 19965963
    Crossref | PubMed | ISI | Google Scholar
  • 540 Spina RJ, Chi MMY, Hopkins MG, Nemeth PM, Lowry OH, Holloszy JO Mitochondrial enzymes increase in muscle in response to 7–10 days of cycle exercise. J Appl Physiol 80 199622502254
    Link | ISI | Google Scholar
  • 541 Staal FJT, Anderson MT, Staal GEJ, Herzenberg LA, Gitler C, Herzenberg LA Redox regulation of signal transduction: tyrosine phosphorylation of calcium influx. Proc Natl Acad Sci USA 91 199436193622
    Crossref | PubMed | ISI | Google Scholar
  • 542 Staal FJT, Roederer M, Herzenberg LA, Herzenberg LA Intracellular thiols regulate activation of nuclear factor κB and transcription of human immunodeficiency virus. Proc Natl Acad Sci USA 87 199099439947
    Crossref | PubMed | ISI | Google Scholar
  • 543 Stadtman ER Oxidation of proteins by mixed-function oxidation systems: implication in protein turnover, aging and neutrophil function. Trends Biochem Sci 11 19861112
    Crossref | ISI | Google Scholar
  • 544 Stadtman ER Protein oxidation and aging. Science 257 199212201224
    Crossref | PubMed | ISI | Google Scholar
  • 545 Stadtman ER Oxidation of free amino acids and amino acid residues in proteins by radiolysis and by metal-catalyzed reactions. Annu Rev Biochem 62 1993797821
    Crossref | PubMed | ISI | Google Scholar
  • 546 Stamler JS, Single D, Loscalzo J Biochemistry of nitric oxide and its redox-activated forms. Science 258 199218981902
    Crossref | PubMed | ISI | Google Scholar
  • 547 Stampfer MJ, Hennekens CH, Manson JE, Colditz GA, Rosner B, Willett WC Vitamin E consumption and the risk of coronary disease in women. N Engl J Med 328 199314441449
    Crossref | PubMed | ISI | Google Scholar
  • 548 Stein B, Rahmsdorf HJ, Steffen A, Litfin M, Herrlich P UV-induced DNA damage is an intermediate step in UV-induced expression of human immunodeficiency virus type 1, collagenase, c-fos, and metallothionein. Mol Cell Biol 9 198951695181
    Crossref | PubMed | ISI | Google Scholar
  • 549 Steinbeck MJ, Khan AU, Karnovsky MJ Extracellular production of singlet oxygen by stimulated macrophages quantified using 9,10-diphenylanthracene and perylene in a polystyrene film. J Biol Chem 268 19931564915654
    PubMed | ISI | Google Scholar
  • 550 Steinberg D, Parthasarathy S, Carew TE, Khoo JC, Witztum JL Beyond cholesterol: modifications of low-density lipoprotein that increase its atherogenicity. N Engl J Med 320 1989915920
    Crossref | PubMed | ISI | Google Scholar
  • 551 Storz G, Tartaglia LA, Ames BN Transcriptional regulator of oxidative stress-inducible genes: direct activation by oxidation. Science 248 1990189194
    Crossref | PubMed | ISI | Google Scholar
  • 552 Su B, Jacinto E, Hibi M, Kallunki T, Karin M, Ben-Neriah Y JNK is involved in signal integration during costimulation of T lymphocytes. Cell 77 1994727736
    Crossref | PubMed | ISI | Google Scholar
  • 553 Suh YA, Arnold RS, Lassegue B, Shi J, Xu X, Sorescu D, Chung AB, Griendling KK, Lambeth JD Cell transformation by the superoxide-generating oxidase Mox1. Nature 401 19997982
    Crossref | PubMed | ISI | Google Scholar
  • 554 Sulciner DJ, Irani K, Yu ZX, Ferrans VJ, Goldschmidt-Clermont P, Finkel T Rac1 regulates a cytokine-stimulated, redox-dependent pathway necessary for NFκB activation. Mol Cell Biol 16 199671157121
    Crossref | PubMed | ISI | Google Scholar
  • 555 Sullivan SG, Chiu DT, Errasfa M, Wang JM, Qi JS, Stern A Effects of H2O2 on protein tyrosine phosphatase activity in HER14 cells. Free Radical Biol Med 16 1994399403
    Crossref | PubMed | ISI | Google Scholar
  • 556 Sun H, Charles CH, Lau LF, Tonks NK MKP-1 (3CH134), an immediate early gene product, is a dual specificity phosphatase that dephosphorylates MAP kinase in vivo. Cell 75 1993487493
    Crossref | PubMed | ISI | Google Scholar
  • 557 Sundaresan M, Zu-Xi Y, Ferrans VJ, Irani K, Finkel T Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science 270 1995296299
    Crossref | PubMed | ISI | Google Scholar
  • 558 Suzukawa K, Miura K, Mitsushita J, Resau J, Hirose K, Crystal R, Kamata T Nerve growth factor-induced neuronal differentiation requires generation of Rac1-regulated reactive oxygen species. J Biol Chem 275 20001317513178
    Crossref | PubMed | ISI | Google Scholar
  • 559 Suzuki YJ, Ford GD Redox regulation of signal transduction in cardiac and smooth muscle. J Mol Cell Cardiol 31 1999345353
    Crossref | PubMed | ISI | Google Scholar
  • 560 Suzuki YJ, Mizuno M, Packer L Transient overexpression of catalase does not inhibit TNF- or PMA-induced NF-κB activation. Biochem Biophys Res Commun 210 1995537541
    Crossref | PubMed | ISI | Google Scholar
  • 561 Suzuki YJ, Packer L Inhibition of NF-κB activation by vitamin E derivatives. Biochem Biophys Res Commun 193 1993277283
    Crossref | PubMed | ISI | Google Scholar
  • 562 Szatrowski TP, Nathan CE Production of large amounts of hydrogen peroxide by human tumor cells. Cancer Res 51 1991794798
    PubMed | ISI | Google Scholar
  • 563 Tak PP, Zvaifler NJ, Green DR, Firestein GS Rheumatoid arthritis and p53: how oxidative stress might alter the course of inflammatory diseases. Immunol Today 21 20007881
    Crossref | PubMed | Google Scholar
  • 564 Tamura S, Brown TA, Whipple JH, Fujita-Yamaguchi Y, Dubler RE, Cheng K, Larner J A novel mechanism for the insulin-like effect of vanadate on glycogen synthase in rat adipocytes. J Biol Chem 259 198466506658
    PubMed | ISI | Google Scholar
  • 565 Taniguchi Y, Taniguchi UY, Mori K, Yodoi J A novel promoter sequence is involved in the oxidative stress-induced expression of the adult T cell leukemia-derived factor ADF/human thioredoxin (Trx) gene. Nucleic Acids Res 24 199627462752
    Crossref | PubMed | ISI | Google Scholar
  • 566 Taub J, Lau JF, Ma C, Hahn JH, Hoque R, Rothblatt J, Chalfie M A cytosolic catalase is needed to extend adult lifespan in C. elegans darf-C and clk-1 mutants. Nature 399 1999162166
    Crossref | PubMed | ISI | Google Scholar
  • 567 Tauber AI, Borregaard N, Simons E, Wright J Chronic granulomatous disease: a syndrome of phagocyte oxidase deficiencies. Medicine 62 1983286309
    Crossref | PubMed | ISI | Google Scholar
  • 568 Tayek JA A review of cancer cachexia and abnormal glucose metabolism in humans with cancer. J Am Coll Nutr 11 1992445456
    Crossref | PubMed | ISI | Google Scholar
  • 569 Tenhaken R, Levine A, Brisson LF, Dixon RA, Lamb C Function of the oxidative burst in hypersensitive disease resistance. Proc Natl Acad Sci USA 92 199541584163
    Crossref | PubMed | ISI | Google Scholar
  • 570 Terasawa Y, Ladha Z, Leonard SW, Morrow JD, Newland D, Sanan D, Packer L, Traber MG, Farese RV Increased atherosclerosis in hyperlipidemic mice deficient in α-tocopherol transfer protein and vitamin E. Proc Natl Acad Sci USA 97 20001383013834
    Crossref | PubMed | ISI | Google Scholar
  • 571 Thannickal VJ, Day RM, Klinz SG, Bastien MC, Larios JM, Fanburg BL Ras-dependent and -independent regulation of reactive oxygen species by mitogenic growth factors and TGF-β1. FASEB J 14 200017411748
    Crossref | PubMed | ISI | Google Scholar
  • 572 Thannickal VJ, Fanburg BL Activation of an H2O2-generating NADH oxidase in human lung fibroblasts by transforming growth factor beta 1. J Biol Chem 270 19953033430338
    Crossref | PubMed | ISI | Google Scholar
  • 573 Thiagarajan RR, Winn RK, Harlan JM The role of leukocyte and endothelial adhesion molecules in ischemia-reperfusion injury. Thromb Hemostasis 78 1997310314
    Crossref | PubMed | ISI | Google Scholar
  • 574 Thiery J, Teupser D, Walli AK, Ivandic B, Nebendahl K, Stein O, Stein Y, Seidel D Study of causes underlying the low atherosclerotic response to dietary hypercholesterolemia in a selected strain of rabbits. Atherosclerosis 121 19966373
    Crossref | PubMed | ISI | Google Scholar
  • 575 Thippeswamy T, Morris R Cyclic guanosine 3′,5′-monophosphate-mediated neuroprotection by nitric oxide in dissociated cultures of rat dorsal root ganglion neurones. Brain Res 774 1997116122
    Crossref | PubMed | ISI | Google Scholar
  • 576 Tirosh A, Potashnik R, Bashan N, Rudich A Oxidative stress disrupts insulin-induced cellular redistribution of insulin receptor substrate-1 and phosphatidylinositol 3-kinase in 3T3–L1 adipocytes. J Biol Chem 274 19991059510602
    Crossref | PubMed | ISI | Google Scholar
  • 577 Toledano MB, Leonard WJ Modulation of transcription factor NFκB binding activity by oxidation-reduction in vitro. Proc Natl Acad Sci USA 88 199143284332
    Crossref | PubMed | ISI | Google Scholar
  • 578 Toledano MB, Kullik I, Trinh F, Baird PT, Schneider TD, Storz G Redox-dependent shift of OxyR-DNA contacts along an extended DNA-binding site: a mechanism for differential promoter selection. Cell 78 1994897909
    Crossref | PubMed | ISI | Google Scholar
  • 579 Tosaki A, Bagchi D, Hellegouarch A, Pali T, Cordis GA, Das DK Comparisons of ESR and HPLC methods for the detection of hydroxyl radicals in ischemic/reperfused hearts. A relationship between the genesis of oxygen-free radicals and reperfusion-induced arrhythmias. Biochem Pharmacol 45 1993961969
    Crossref | PubMed | ISI | Google Scholar
  • 580 Traencker EB, Wilk S, Baeuerle PA A proteasome inhibitor prevents activation of NF-κB and stabilizes a newly phosphorylated form of IκBα that is still bound to NF-κB. EMBO J 13 199454335441
    Crossref | PubMed | ISI | Google Scholar
  • 581 Tsaneva IR, Weiss B soxR, a locus governing a superoxide response regulon in Escherichia coli K-12. J Bacteriol 172 199041974205
    Crossref | PubMed | ISI | Google Scholar
  • 582 Tu P, Gurney ME, Julien JP, Lee VMY, Trojanowski JQ Oxidative stress, mutant SOD1, and neurofilament pathology in transgenic mouse models of human motor neuron disease. Lab Invest 76 1997441456
    PubMed | ISI | Google Scholar
  • 583 Tucker KR, Seider MJ, Booth FW Protein synthesis rates in atrophied gastrocnemius muscles after limb immobilization. J Appl Physiol 51 19817377
    Link | ISI | Google Scholar
  • 584 Turrens JF, Alexandre A, Lehninger AL Ubisemiquinone is the election donor for superoxide formation by complex III of heart mitochondria. Arch Biochem Biophys 237 1985408414
    Crossref | PubMed | ISI | Google Scholar
  • 585 Tyrrell R Redox regulation and oxidant activation of heme oxygenase-1. Free Radical Res 31 1999335340
    Crossref | PubMed | ISI | Google Scholar
  • 586 Tyrrell RM, Applegate LA, Tromvoukis Y The proximal promoter region of the human heme oxygenase gene contains elements involved in stimulation of transcriptional activity by a variety of agents including oxidants. Carcinogenesis 14 1993761765
    Crossref | PubMed | ISI | Google Scholar
  • 587 Uknes S, Mauch-Mani B, Moyer M, Potter S, Williams S, Dincher S, Chandler D, Slusarenko A, Ward E, Ryals J Acquired resistance in Arabidopsis. Plant Cell 4 1992645656
    Crossref | PubMed | ISI | Google Scholar
  • 588 Um HD, Orenstein JM, Wahl SM Fas mediates apoptosis in human monocytes by a reactive oxygen intermediate dependent pathway. J Immunol 156 199634693477
    PubMed | ISI | Google Scholar
  • 589 Umansky V, Rocha M, Breitkreutz R, Hehner S, Bucur M, Erbe N, Dröge W, Ushmorov A Glutathione is a factor of resistance of Jurkat leukemia cells to nitric oxide-mediated apoptosis. J Cell Biochem 78 2000578587
    Crossref | PubMed | ISI | Google Scholar
  • 590 Ushio-Fukai M, Alexander RW, Akers M, Griendling KK p38 MAP kinase is a critical component of the redox-sensitive signaling pathways by angiotensin II: role in vascular smooth muscle cell hypertrophy. J Biol Chem 273 19981502215029
    Crossref | PubMed | ISI | Google Scholar
  • 591 Ushio-Fukai M, Griendling KK, Becker PL, Alexander RW Role of reactive oxygen species in angiotensin II-induced transactivation of epidermal growth factor receptor in vascular smooth muscle cells. Circulation 100 Suppl 1999I-263
    ISI | Google Scholar
  • 592 Ushmorov A, Ratter F, Lehmann V, Dröge W, Schirrmacher V, Umansky V Nitric oxide-induced apoptosis in human leukemic lines requires mitochondrial lipid degradation and cytochrome c release. Blood 93 199923422352
    PubMed | ISI | Google Scholar
  • 593 Van Dam PS, van Asbeck BS, Erkelens DW, Marx JJM, Gispen WH, Bravenboer B The role of oxidative stress in neuropathy and other diabetic complication. Diabetes Metab Rev 11 1995181192
    Crossref | PubMed | ISI | Google Scholar
  • 594 Van Eden W, Thole JER, van der Zee R, Noordzij A, van Embden JDA, Hensen EJ, Cohen IR Cloning of the mycobacterial epitope recognized by T lymphocytes in adjuvant arthritis. Nature 331 1988171173
    Crossref | PubMed | ISI | Google Scholar
  • 595 Veale DJ, Maple C Cell adhesion molecules in rheumatoid arthritis. Drugs Aging 9 19968792
    Crossref | PubMed | ISI | Google Scholar
  • 596 Vertechy M, Cooper MB, Ghirardi O, Ramacci MT Antioxidant enzyme activities in heart and skeletal muscle of rats of different ages. Exp Gerontol 24 1998211218
    Crossref | ISI | Google Scholar
  • 597 Vestergaard H, Andersen PH, Lund S, Schmitz O, Junker S, Pedersen O Pre- and posttranslational upregulation of muscle-specific glycogen synthase in athletes. Am J Physiol Endocrinol Metab 266 1994E92E101
    Link | ISI | Google Scholar
  • 598 Vestergaard H, Lund S, Larsen FS, Bjerrum OJ, Pedersen O Glycogen synthase and phosphofructokinase protein and mRNA levels in skeletal muscle from insulin-resistant patients with non-insulin-dependent diabetes mellitus. J Clin Invest 91 199323422350
    Crossref | PubMed | ISI | Google Scholar
  • 599 Vina J, Gimeno A, Sastre J, Desco C, Asensi M, Pallardó FV, Cuesta A, Ferrero JA, Terada LS, and Repine JE. Mechanism of free radical production in exhaustive exercise in humans and rats: role of xanthine oxidase and protection by allopurinol. Life. In press.
    Google Scholar
  • 600 Vita JA, Frei B, Holbrook M, Gokce N, Leaf C, Keaney JF l-2-Oxothiazolidine-4-carboxylic acid reverses endothelial dysfunctions in patients with coronary artery disease. J Clin Invest 101 199814081414
    Crossref | PubMed | ISI | Google Scholar
  • 601 Von Zglinicki T, Brunk UT Intercellular interactions under oxidative stress and aging: a hypothesis. Z Gerontol 26 1993215220
    PubMed | ISI | Google Scholar
  • 602 Von Zglinicki T, Nilsson E, Döcke WD, Brunk UT Lipofuscin accumulation and aging of fibroblasts. Gerontology 41 Suppl 2 199595109
    Crossref | PubMed | ISI | Google Scholar
  • 603 Von Zglinicki T, Saretzki G, Döcke W, Lotze C Mild hyperoxia shortens telomeres and inhibits proliferation of fibroblasts: a model for senescence? Exp Cell Res 220 1995186193
    Crossref | PubMed | ISI | Google Scholar
  • 604 Walford RL, Jawaid SW, Naeim F Evidence for in vitro senescence of T lymphocytes cultured from normal human peripheral blood. Age 4 19876771
    Crossref | Google Scholar
  • 605 Wanagat J, Cao Z, Pathare P, Aiken JM Mitochondrial DNA deletion mutations colocalize with segmental electron transport system abnormalities, muscle fiber atrophy, fiber splitting, and oxidative damage in sarcopenia. FASEB J 15 2001322332
    Crossref | PubMed | ISI | Google Scholar
  • 606 Wang CY, Mayo MW, Baldwin ASJ TNF- and cancer therapy-induced apoptosis: potentiation by inhibition of NFκB. Science 274 1996784787
    Crossref | PubMed | ISI | Google Scholar
  • 607 Wang D, Yu X, Cohen RA, Brecher P Distinct effects of N-acetylcysteine and nitric oxide on angiotensin II-induced epidermal growth factor receptor phosphorylation and intracellular Ca2+ levels. J Biol Chem 275 20001222312230
    Crossref | PubMed | ISI | Google Scholar
  • 608 Wang GL, Jiang BH, Rue EA, Semenza GL Hypoxia-inducible factor 1 is a basic helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci USA 92 199555105514
    Crossref | PubMed | ISI | Google Scholar
  • 609 Wang T, Marquardt C, Foker J Aerobic glycolysis during lymphocyte proliferation. Nature 261 1976702705
    Crossref | PubMed | ISI | Google Scholar
  • 610 Wang ZG, Delva L, Gaboli M, Rivi R, Giorgio M, Cordon-Cardo C, Grosveld F, Pandolfi PP Role of PML in cell growth and the retinoic acid pathway. Science 279 199815471551
    Crossref | PubMed | ISI | Google Scholar
  • 611 Warburg O, Posener K, Negelein E über den Stoffwechsel der Carcinomzelle. Biochem Z 152 1924309344
    Google Scholar
  • 612 Wei L, Hubbard SR, Hendrickson WA, Ellis L Expression, characterization, and crystallization of the catalytic core of the human insulin receptor protein-tyrosine kinase domain. J Biol Chem 270 199581228130
    Crossref | PubMed | ISI | Google Scholar
  • 613 Weindruch R, Sohal RS Seminars in medicine of the Beth Israel Deaconess Medical Center. Caloric intake and aging. N Engl J Med 337 1997986994
    Crossref | PubMed | ISI | Google Scholar
  • 614 Weindruch R, Walford RL, Fligiel S, Guthrie D The retardation of aging by dietary restriction in mice: longevity, cancer, immunity and lifetime energy intake. J Nutr 116 1986641654
    Crossref | PubMed | ISI | Google Scholar
  • 615 White AA, Crawford KM, Patt CS, Lad PJ Activation of soluble guanylate cyclase from rat lung by incubation or by hydrogen peroxide. J Biol Chem 251 197673047312
    PubMed | ISI | Google Scholar
  • 616 White MF, Shoelson SE, Keutmann H, Kahn CR A cascade of tyrosine autophosphorylation in the β-subunit activates the phosphotransferase of the insulin receptor. J Biol Chem 263 198829692980
    Crossref | PubMed | ISI | Google Scholar
  • 617 White RF Salicylic acid and its derivatives induce gene expression. Virology 99 1978410420
    Crossref | ISI | Google Scholar
  • 618 Wick G, Romen M, Amberger A, Metzler B, Mayr M, Falkensammer G, Xu Q Atherosclerosis, autoimmunity, and vascular-associated lymphoid tissue. FASEB J 11 199711991207
    Crossref | PubMed | ISI | Google Scholar
  • 619 Wick G, Schett G, Amberger A, Kleindienst R, Xu Q Is atherosclerosis an immunologically mediated disease? Immunol Today 16 19952733
    Crossref | PubMed | Google Scholar
  • 620 Wilden PA, Pessin JE Differential sensitivity of the insulin-receptor kinase to thiol and oxidizing agents in the absence and presence of insulin. Biochem J 245 1987325331
    Crossref | PubMed | ISI | Google Scholar
  • 621 Wilhelm D, Bender K, Knebel A, Angel P The level of intracellular glutathione is a key regulator for the induction of stress-activated signal transduction pathways including Jun N-terminal protein kinases and p38 kinase by alkylating agents. Mol Cell Biol 17 199747924803
    Crossref | PubMed | ISI | Google Scholar
  • 622 Williams JC, Weijland A, Gonfloni S, Thompson A, Courtneidge SA, Superti-Furga G, Wierenga RK The 2.35 Å crystal structure of the inactivated form of chicken Src: a dynamic molecule with multiple regulatory interactions. J Mol Biol 274 1997757775
    Crossref | PubMed | ISI | Google Scholar
  • 623 Williams MS, Henkart PA Role of reactive oxygen intermediates in TCR-induced death of T cell blasts and hybridomas. J Immunol 157 199623952402
    PubMed | ISI | Google Scholar
  • 624 Winge P, Brembu T, Bones AM Cloning and characterization of rac-like cDNAs from Arabidopsis thaliana. Plant Mol Biol 35 1997483495
    Crossref | PubMed | ISI | Google Scholar
  • 625 Witkowski JA Cell ageing in vitro: a historical perspective. Exp Gerontol 22 1987231248
    Crossref | PubMed | ISI | Google Scholar
  • 626 Wohaieb SA, Godin DV Alterations in free radical tissue-defence mechanisms in streptozotocin-induced diabetes in rat. Effects of insulin treatment. Diabetes 36 198710141018
    Crossref | PubMed | ISI | Google Scholar
  • 627 Wolbach SB, Howe PR Tissue changes following deprivation of fat-soluble A vitamin. J Exp Med 42 1925753777
    Crossref | PubMed | Google Scholar
  • 628 Wolf SF, Jiang ZY, Hunt JV Protein glycation and oxidative stress in diabetes mellitus and ageing. Free Radical Biol Med 10 1991339352
    Crossref | PubMed | ISI | Google Scholar
  • 629 Wolff SP Diabetes mellitus and free radicals. Free radicals, transition metals and oxidative stress in the aetiology of diabetes mellitus and complications. Br Med Bull 49 1993642652
    Crossref | PubMed | ISI | Google Scholar
  • 630 Wolff SP, Crabbe MJC, Thornalley PJ The autooxidation of simple monosaccharides. Experientia 40 1984244246
    Crossref | Google Scholar
  • 631 Wolff SP, Garner A, Dean RT Free radicals, lipids, and protein degradation. Trends Biochem Sci 11 19862731
    Crossref | ISI | Google Scholar
  • 632 Wolin MS, Burke-Wolin TM, Mohazzab-H KM Roles of NAD(P)H oxidases and reactive oxygen species in vascular oxygen sensing mechanisms. Respir Physiol 115 1999229238
    Crossref | PubMed | Google Scholar
  • 633 Wu G, Shortt BJ, Lawrence EB, León J, Fitzsimmons KC, Levine EB, Raskin I, Shah DM Activation of host defense mechanisms by elevated production of H2O2 in transgenic plants. Plant Physiol 115 1997427435
    Crossref | PubMed | ISI | Google Scholar
  • 634 Wu J, Dunahm WR, Weiss B Overproduction and physical characterization of SoxR, a [2Fe-2S] protein that governs an oxidative response regulon in Escherichia coli. J Biol Chem 270 19951032310327
    Crossref | PubMed | ISI | Google Scholar
  • 635 Wu M, Lee H, Bellas RE, Schauer SL, Arsura M, Katz D, FitzGerald MJ, Rothstein TL, Sherr DH, Sonenshein GE Inhibition of NFκB B/Re1 induces apoptosis of murine B cells. EMBO J 15 199646824690
    Crossref | PubMed | ISI | Google Scholar
  • 636 Wyllie AH, Kerr JFR, Currie AE Cell death: the significance of apoptosis. Int Rev Cytol 68 1980251306
    Crossref | PubMed | Google Scholar
  • 637 Xanthoudakis S, Curran T Identification and characterization of Ref-1, a nuclear protein that facilitates AP-1 DNA binding activity. EMBO J 11 1992653665
    Crossref | PubMed | ISI | Google Scholar
  • 638 Xanthoudakis S, Miao G, Wang F, Pan YC, Curran T Redox activation of Fos-Jun DNA binding activity is mediated by a DNA repair enzyme. EMBO J 11 199233233335
    Crossref | PubMed | ISI | Google Scholar
  • 639 Xie QW, Cho HJ, Calaycay J, Mumford RA, Swiderek KM, Lee TD, Ding A, Troso T, Nathan C Cloning and characterization of inducible nitric oxide synthase from mouse macrophages. Science 256 1992225228
    Crossref | PubMed | ISI | Google Scholar
  • 640 Xiong Y, Hannon GJ, Zhang H, Casso D, Kobayashi R, Beach D p21 is a universal inhibitor of cyclin kinases. Nature 366 1993701704
    Crossref | PubMed | ISI | Google Scholar
  • 641 Xu Q, Kleindienst R, Waitz W, Dietrich H, Wick G Increased expression of heat shock protein 65 coincides with a population of infiltrating T lymphocytes in atherosclerotic lesions of rabbits specifically responding to heat shock protein 65. J Clin Invest 91 199326932702
    Crossref | PubMed | ISI | Google Scholar
  • 642 Xu W, Harrison SC, Eck MJ Three-dimensional structure of the tyrosine kinase c-Src. Nature 385 1997595602
    Crossref | PubMed | ISI | Google Scholar
  • 643 Yamaguchi H, Hendrickson WA Structural basis for activation of human lymphocyte kinase Lck upon tyrosine phosphorylation. Nature 384 1996484489
    Crossref | PubMed | ISI | Google Scholar
  • 644 Yamamoto S Mammalian lipoxygenases: molecular structures and functions. Biochim Biophys Acta 1128 1992117131
    Crossref | PubMed | ISI | Google Scholar
  • 645 Yan SD, Schmidt AM, Anderson GM, Zhang J, Brett J, Zou YS, Pinsky D, Stern D Enhanced cellular oxidative stress by the interaction of advanced glycation end products with their receptors/binding protein. J Biol Chem 269 199498899897
    PubMed | ISI | Google Scholar
  • 646 Yarom R, Sagher U, Havivi Y, Peled IJ, Wexler MR Myofibers in tongues of Down's syndrome. J Neurol Sci 73 1986279287
    Crossref | PubMed | ISI | Google Scholar
  • 647 Yarom R, Sherman Y, Sagher U, Peled IJ, Wexler MR Elevated concentrations of elements and abnormalities of neuromuscular junctions in tongue muscles of Down's syndrome. J Neurol Sci 79 1987315326
    Crossref | PubMed | ISI | Google Scholar
  • 648 Yarom R, Sapoznikov D, Havivi Y, Avraham KB, Schickler M, Groner Y Premature aging changes in neuromuscular junctions of transgenic mice with an extra human CnZnSOD gene: a model for tongue pathology in Down's syndrome. J Neurol Sci 88 19884153
    Crossref | PubMed | ISI | Google Scholar
  • 649 Yashpe-Purer Henis Y, Yashpe J Regulation of catalase level in Escherichia coli K12. J Microbiol 23 19778491
    ISI | Google Scholar
  • 650 Yki-Järvinen H, Young AA, Lamkin C, Foley JE Kinetics of glucose disposal in whole body and across the forearm in man. J Clin Invest 79 198717131719
    Crossref | PubMed | ISI | Google Scholar
  • 651 Yoshizumi M, Abe J, Haendeler J, Huang Q, Berk BC Src and Cas mediate JNK activation but not ERK1/2 and p38 kinases by reactive oxygen species. J Biol Chem 275 20001170611712
    Crossref | PubMed | ISI | Google Scholar
  • 652 Young T, Palta M, Skatrud J, Weber S, Badr S The occurrence of sleep-disordered breathing among middle-aged adults. N Engl J Med 328 199312301235
    Crossref | PubMed | ISI | Google Scholar
  • 653 Zafari AM, Ushio-Fukai M, Akers M, Yin Q, Shah A, Harrison DG, Taylor WR, Griendling KK Role of NADH/NADPH oxidase-derived H2O2 in angiotensin II-induced vascular hypertrophy. Hypertension 32 1998488495
    Crossref | PubMed | ISI | Google Scholar
  • 654 Zainal TA, Oberley TD, Allison DB, Szweda LI, Weindruch R Caloric restriction of rhesus monkeys lowers oxidative damage in skeletal muscle. FASEB J 14 200018251836
    Crossref | PubMed | ISI | Google Scholar
  • 655 Zamzami Marchetti NP, Castedo M, Decaudin D, Macho A, Hirsch T, Susin SA, Petit PX, Mignotte B, Kroemer G Sequential reduction of mitochondrial transmembrane potential and generation of reactive oxygen species in early programmed cell death. J Exp Med 182 1995367377
    Crossref | PubMed | ISI | Google Scholar
  • 656 Zheng J, Knighton DR, Xuong NH, Taylor SS, Sowadski JM, Ten EL Crystal structures of the myristoylated catalytic subunit of cAMP-dependent protein kinase reveal open and closed conformations. Protein Sci 2 199315591573
    Crossref | PubMed | ISI | Google Scholar
  • 657 Zheng M, Åslund F, Storz G Activation of the OxyR transcription factor by reversible disulfide bond formation. Science 279 199817181721
    Crossref | PubMed | ISI | Google Scholar
  • 658 Zheng M, Storz G Redox sensing by prokaryotic transcription factors. Biochem Pharmacol 59 200016
    Crossref | PubMed | ISI | Google Scholar
  • 659 Zhu H, Bunn HF Oxygen sensing and signaling: impact on the regulation of physiologically important genes. Respir Physiol 115 1999239247
    Crossref | PubMed | Google Scholar
  • 660 Zweier JL, Broderick R, Kuppusamy P, Thompson-Gorman S, Lutty GA Determination of the mechanism of free radical generation in human aortic endothelial cells exposed to anoxia and reoxygenation. J Biol Chem 269 19942415624162
    PubMed | ISI | Google Scholar

Source: https://journals.physiology.org/doi/full/10.1152/physrev.00018.2001

Posted by: sheridansheridanerpeldinge0301595.blogspot.com