Acessibilidade / Reportar erro

Superoxide: a two-edged sword

Abstract

Superoxide (O2-) is the compound obtained when oxygen is reduced by one electron. For a molecule with an unpaired electron, O2- is surprisingly inert, its chief reaction being a dismutation in which it reacts with itself to form H2O2 and oxygen. The involvement of O2- in biological systems was first revealed by the discovery in 1969 of superoxide dismutase, an enzyme that catalyzes the dismutation of O2-. Since then it has been found that biological systems produce a bewildering variety of reactive oxidants, all but a few arising ultimately from O2-. These oxidants include O2- itself, H2O2 and alkyl peroxides, hydroxyl radical and other reactive oxidizing radicals, oxidized halogens and halamines, singlet oxygen, and peroxynitrite. These various oxidants are able to damage molecules in their environment, and are therefore very dangerous. They are thought to participate in the pathogenesis of a number of common diseases, including among others malignancy, by their ability to mutate the genome, and atherosclerosis, by their capacity for oxidizing lipoproteins. Their properties are put to good use, however, in host defense, where they serve as microbicidal and parasiticidal agents, and in biological signalling, where their liberation in small quantities results in redox-mediated changes in the functions of enzymes and other proteins

NADPH oxidase; Superoxide; Oxidative stress; Antioxidant; Regulation; Host defense


Braz J Med Biol Res, February 1997, Volume 30(2) 141-155

Superoxide: a two-edged sword

B.M. Babior

The Scripps Research Institute, La Jolla, CA 92037, USA

Text

References

Correspondence and Footnotes Correspondence and Footnotes Correspondence and Footnotes

Abstract

Superoxide (O2-) is the compound obtained when oxygen is reduced by one electron. For a molecule with an unpaired electron, O2- is surprisingly inert, its chief reaction being a dismutation in which it reacts with itself to form H2O2 and oxygen. The involvement of O2- in biological systems was first revealed by the discovery in 1969 of superoxide dismutase, an enzyme that catalyzes the dismutation of O2-. Since then it has been found that biological systems produce a bewildering variety of reactive oxidants, all but a few arising ultimately from O2-. These oxidants include O2- itself, H2O2 and alkyl peroxides, hydroxyl radical and other reactive oxidizing radicals, oxidized halogens and halamines, singlet oxygen, and peroxynitrite. These various oxidants are able to damage molecules in their environment, and are therefore very dangerous. They are thought to participate in the pathogenesis of a number of common diseases, including among others malignancy, by their ability to mutate the genome, and atherosclerosis, by their capacity for oxidizing lipoproteins. Their properties are put to good use, however, in host defense, where they serve as microbicidal and parasiticidal agents, and in biological signalling, where their liberation in small quantities results in redox-mediated changes in the functions of enzymes and other proteins.

Key words: NADPH oxidase, superoxide, oxidative stress, antioxidant, regulation, host defense

Introduction

The reduction of a molecule of oxygen to two molecules of water is the major source of energy in aerobic biological systems. This reduction requires 4 electrons:

If these electrons are passed to oxygen one at a time, a series of partially reduced products is generated (1). The first of these is superoxide (O2-):

Reduction of O2- by the second electron yields hydrogen peroxide:

·OH (hydroxyl radical) plus the first molecule of water arise when the third electron is passed on to H2O2:

And finally, the fourth electron produces the second molecule of water from the hydroxyl radical:

Unlike most molecules with unpaired electrons, O2- is surprisingly inert. In aqueous systems, its principal reaction is with itself to generate a molecule of H2O2 and a molecule of oxygen in a dismutation reaction (2):

It is also a weak base, its conjugate acid being the much more reactive hydroperoxyl radical:

It does, however, have certain other chemical properties that are important in a biological context. These include 1) its participation in the so-called Haber-Weiss reaction to generate ·OH and in a closely related reaction to generate alkoxyl radical (RO·), reactions that are catalyzed by transition metals such as iron or copper (3-5):

2) its ability to obtain Fe2+ needed for the Haber-Weiss reaction by liberating it from the iron storage protein ferritin and from iron-sulfur proteins such as aconitase (6,7); 3) its reaction with ·OH to form singlet oxygen (8):

and 4) its reaction with nitric oxide to form peroxynitrite, a highly reactive oxidant that breaks down to produce a nitrating agent (9,10):

In addition, the H2O2 produced by the dismutation of O2- is used by phagocytes to oxidize halide ions to the level of hypohalous acids (e.g., HOCl) (11), a group of highly reactive compounds which in turn react with amines to produce halamines (e.g., NH2Cl), some of which are even more reactive than the hypohalous acids (12,13). In turn, the hypohalous acids can react with H2O2 to generate singlet oxygen (14,15). For example,

All this was of little interest to biologists, however, until 1969, when McCord and Fridovich (16) discovered superoxide dismutase. This enzyme destroys O2- by catalyzing the dismutation reaction described above. The ubiquitous occurrence of an enzyme that catalyzes the destruction of O2- implied that O2- had to be participating in an important way in the biological economy (1,17). Certain aspects of the nature of its participation are reviewed here.

Superoxide the evil

As the precursor of a large number of highly reactive oxidizing agents, including oxidizing radicals, singlet oxygen, peroxynitrite and oxidized halogens such as HOCl, O2- clearly had the potential to inflict considerable damage on biological systems. That O2- could realize this potential was shown in numerous experiments, at first indirectly by demonstrations, for example, that oxidative stress induced substantial increases in the superoxide dismutase concentration of E. coli (18), and later directly by genetic experiments with bacteria and eukaryotes containing no superoxide dismutase or excess superoxide dismutase (19-24). Damage to DNA (25), proteins (26) and lipids (27,28) have all been documented as consequences of exposure to O2- and its descendants. DNA damage may lead to the production of abnormal bases such as thymine glycol and 8-hydroxyguanine (25,29-31) or to strand breakage through a series of reactions initiated by the abstraction of a 4' hydrogen atom from a ribose residue (32-35). On proteins, susceptible residues such as cysteine and histidine are oxidized (36-38), leading in some cases to the production of oxo groups that can be assayed to provide an index of oxidative damage to proteins (36,39). In addition, tyrosine residues are nitrated, a consequence of the spontaneous decay of ONOO- into a nitrating agent of some sort (40,41). Hypohalous acids will decarboxylate a-amino acids to aldehydes (11), and will halogenate tyrosine and heterocycles (e.g., adenosine, NAD) (42-44). Polyunsaturated fatty acid residues on phospholipids and triglycerides undergo peroxidation to form toxic alkyl hydroperoxides and aldehydes (45-47) (see Diagram 1).

Among these is malondialdehyde, produced by the oxidation at the two double bonds flanking the methylene carbons of polyunsaturated fatty acids:

Because of its bifunctionality, malondialdehyde is an effective cross-linking agent, able to act on macromolecules such as DNA and proteins. In addition, malondialdehyde reacts with thiobarbituric acid to form a pink compound whose color has been used as a measure of lipid peroxidation. Cholesterol is oxidized at its susceptible 4- and 6-carbons (see Diagram 2).

These oxidation reactions are currently believed to participate in the pathogenesis of a number of important degenerative diseases, including atherosclerosis, in which the uptake of oxidized lipoproteins via the scavenger receptor of macrophages is thought to be an early step in the formation of the endothelial foam cells and lipid deposits characteristic of that condition (48,49); malignancies, some of which may arise as a result of oncogene mutations caused by oxidative damage to DNA (50,51); arthritis, due to joint damage inflicted in part by oxidants released at sites of inflammation (52), and possibly aging itself (53,54), not a disease but an inevitable consequence of living.

Because of the evil nature of these oxidants, many defenses have been erected against them. These defenses include both enzymes and low molecular weight compounds.

Antioxidant enzymes

Four enzymes comprise three enzyme-based antioxidant systems that deal with oxidants formed by the partial reduction of oxygen:

1. Superoxide dismutase (2,55,56). Superoxide dismutase catalyzes the destruction of O2- by converting it to oxygen and H2O2:

The uncatalyzed reaction is very rapid, proceeding with a rate constant of ca. 106-107 M-1s-1 at the pH values prevailing in tissues. Superoxide dismutase, however, greatly accelerates the rate of destruction of O2-, in part by converting a second order reaction to a first order reaction. Because of the effect of superoxide dismutase, steady-state concentrations of O2- in tissues are many orders of magnitude lower than they would be if the elimination of O2- was solely dependent on its spontaneous dismutation.

Two forms of superoxide dismutase are present in eukaryotic cells: a form that contains Cu2+ and Zn2+, the former serving as the redox center and the latter as a structural element, and a form that contains only one metal, namely Mn2+, which functions as the redox center. The Cu2+/Zn2+ form, a 32-kDa dimer, is found in the cytosol, while the Mn2+ form, an 80-kDa tetramer, is located in mitochondria. The Mn2+ form is also found in bacteria, as is a third form of superoxide dismutase containing Fe2+ as its redox element. The concentrations of the Cu2+/Zn2+ and Fe2+ forms of superoxide dismutase are unaffected by oxidative stresses, but the Mn2+ form is inducible in both bacteria and eukaryotic cells, its activity increasing with oxidative stress.

Mutant forms of the Cu2+/Zn2+ enzyme appear to explain the familial forms of a fatal neurological disease known as amyotrophic lateral sclerosis, or motor neuron disease (57). In this condition, the motor neurons in the patient's cerebral cortex and spinal cord degenerate over the course of a few years, leading to weakness and eventually paralysis, with death from pneumonia caused by the inability of the patient to clear respiratory secretions. The mutant enzymes dismute superoxide in a normal fashion, but they have excess peroxidase activity, an activity present in normal Cu2+/Zn2+ dismutase to only a very limited extent (58). It is presently thought that the oxidative damage inflicted by the increased peroxidase activity of the mutant dismutase is responsible for the early death of these neurons.

2. Catalase (59,60). The H2O2 produced by the dismutation of O2- or generated by H2O2-generating oxidases (e.g., D-amino acid oxidase) is handled by two systems: catalase and a glutathione-dependent antioxidant system that reduces H2O2 to water at the expense of NADPH. Catalase is a tetrameric heme enzyme of 240-kDa mass that catalyzes the dismutation of H2O2 to oxygen and water:

Catalase is erroneously said to work only at high concentrations of H2O2, and to serve principally as a backup for the glutathione-dependent system to be discussed below, but the enzyme has a binding site for NADPH, and when this site is occupied, catalase operates at H2O2 concentrations in the vicinity of those at which the glutathione-dependent systems operate (61). It is therefore likely that some half the H2O2 produced in the cell is destroyed by catalase. Catalase deficiency exists, but is relatively innocuous; the Swiss type is asymptomatic, while the Japanese variety is associated only with ulcers of the oral cavity.

3. The glutathione-dependent antioxidant system (62-65). The glutathione-dependent antioxidant system consists of glutathione plus two enzymes: glutathione peroxidase and glutathione reductase. As this system operates, glutathione cycles between its oxidized and reduced forms. The reactions catalyzed by these enzymes are:

Like other enzymes that catalyze the interconversion of sulfhydryl groups and disulfides, the 22-kDa glutathione reductase uses FAD as its cofactor. Glutathione peroxidase, another 22-kDa protein, is unusual, however, in that the redox element in its active site is selenocysteine. The selenocysteine is introduced into the protein by a special t-RNA that is initially charged with serine but undergoes a series of reactions that convert it to t-RNAselenoCys. Selenocysteine is encoded by the triplet UGA, which ordinarily introduces a stop but in the context of the glutathione peroxidase mRNA is recognized by the selenocysteine-linked t-RNA (66). The antioxidant properties of selenium are explained by its occurrence in glutathione peroxidase.

Families with inherited deficiencies of glutathione peroxidase (67,68) and glutathione reductase (69) have been reported. Affected members manifest a mild to moderately severe hemolytic anemia that is aggravated by infection and by oxidant drugs such as nitrofurantoin and certain sulfonamides. Selenium deficiency on a nutritional basis leads to a cardiomyopathy that may in part represent oxidative damage due to glutathione peroxidase deficiency but, in addition, is likely to reflect injury caused by the deficiency of other selenium-containing enzymes, because it is not seen in inherited glutathione peroxidase deficiency (70,71).

Lipid hydroperoxides, which are formed during the peroxidation of lipids containing unsaturated fatty acids, are reduced, not by the usual glutathione peroxidase, but by a special enzyme designed specifically to handle peroxidized fatty acids in phospholipids. This enzyme, known as phospholipid hydroperoxide glutathione peroxidase (72), is an 18-kDa protein that can reduce both H2O2 and lipid hydroperoxides to the corresponding hydroxides (water and a lipid hydroxide, respectively). In contrast to the phospholipid hydroperoxide glutathione peroxidase, ordinary glutathione peroxidase is unable to act on lipid hydroperoxides.

Low molecular weight antioxidants

Many low molecular weight compounds can act as biological antioxidants, including among others carotenoids, bilirubin and uric acid. The most important of these, however, are two vitamins: ascorbic acid (vitamin C) (73-76), and a-tocopherol (vitamin E) (77-80). Ascorbic acid, a very water soluble compound, reacts with free radicals that arise in the aqueous compartments of tissues, forming the innocuous ascorbate semiquinone (81,82) (see Diagram 3).

The semiquinone is consumed in a dismutation reaction in which two semiquinone molecules react to produce a molecule of ascorbate and a molecule of dehydroascorbate (see Diagram 4).

The dehydroascorbate is then enzymatically reduced back to ascorbate by dehydroascorbate reductase (83).

a-Tocopherol, a highly lipophilic molecule, is the chief antioxidant in biological membranes. It reacts with free radicals to form the highly stable tocopherol semiquinone (84) (see Diagram 5).

The semiquinone is then reduced back to the alcohol by ascorbic acid (85-87).

A small antioxidant with special properties is taurine (2-aminoethyl sulfonic acid). This compound reacts with hypohalous acids and halamines (77-80,88-91), which are routinely generated by phagocytes for use as microbicidal killing agents (see below, Superoxide the good) but whose great reactivity allows them to inflict major damage in the tissues in which they are released. The product of the reaction between an oxidized halogen and taurine, however, is a halamine with exceptionally low reactivity for this class of compounds - such low reactivity, in fact, that it is harmless to tissues (92). An example is the reaction between taurine and hypochlorous acid to form taurine chloramine:

Taurine is therefore able to detoxify these very reactive and dangerous oxidized halogens by converting them to innocuous compounds.

Superoxide the good

O 2 - production by phagocytes

The consequences of O2- production in tissues are not all bad. In fact, the production of O2- can be lifesaving. This was first demonstrated by the surprising finding that O2- is produced in large quantities by stimulated phagocytes (93,94), and that individuals with chronic granulomatous disease, an inherited disorder in which phagocytes are unable to manufacture O2-, are highly susceptible to very dangerous bacterial and fungal infections that formerly killed most of the patients before they reached their tenth birthday (95), though recently the advent of newer modes of treatment has greatly improved their outlook (96). The O2- produced by these cells is made by the leukocyte NADPH oxidase (97), a membrane-associated enzyme, that catalyzes the one-electron reduction of oxygen at the expense of NADPH:

The oxidase is dormant in resting cells, but develops catalytic activity when the cells encounter a microorganism or are exposed to any of several soluble stimuli, including N-formylated peptides, the complement polypeptide C5a or leukotriene B4.

The O2- produced by these cells is only weakly microbicidal, though it can inactivate bacterial iron-sulfur proteins such as aconitase. The major microbicidal oxidant of phagocytes is HOCl, produced by the H2O2generated by the dismutation of O2- (98-100). HOCl production is catalyzed by myeloperoxidase, which is abundant in neutrophils and monocytes and catalyzes the two-electron oxidation of the chloride ion by hydrogen peroxide (12,101):

Myeloperoxidase can also oxidize Br- and I- to the corresponding hypohalous acids (102). An enzyme in eosinophils, the eosinophil peroxidase, catalyzes the same reaction, except that Cl- is not a substrate; the chief product of the eosinophil peroxidase reaction is HOBr (102,103). The hypohalous acid will react with any of the hundreds of amines present in the cell to form a vast array of halamines (12,88,102,104) whose toxicities range from none to extreme, the degree of toxicity roughly correlating with the lipid solubility of the halamine.

Recent work with E. coli has revealed a mechanism by which oxidized halogens can kill a microorganism. It has been observed that the lethal action of HOCl against E. coli takes place on the membrane. This membrane contains a binding site to which the origin of replication of the E. coli genome (oriC) has to attach before the DNA can be copied during bacterial replication. HOCl is able to inactivate this binding site, and experiments have shown that the fraction of bacteria killed by HOCl at a given time is virtually identical to the fraction of binding sites that have been inactivated (105). These results imply that the destruction of this binding site is tantamount to the destruction of the microorganism itself.

The oxidizing radicals produced from the O2- in the Haber-Weiss and related reactions also participate in the oxidative killing of microorganisms, but principally as a backup system. This is shown by the observation that patients with myeloperoxidase deficiency have little or no problem with infections (106,107), in contrast to patients with chronic granulomatous disease, whose very high susceptibility to such infections was discussed above. This indicates the occurrence in phagocytes of a backup microbicidal system that is dependent on oxygen and is active in patients whose neutrophils and monocytes are unable to manufacture oxidized halogens (i.e., cells deficient in myeloperoxidase). This system is highly likely to employ the oxidizing radicals known to be produced by these phagocytes.

A possible antioxidant activity of O 2 - itself

There appears to be an optimum for the intracellular concentration of superoxide dismutase. Too much superoxide dismutase can be just as harmful as too little (108-111). This finding has given rise to the speculation that a little intracellular O2- is necessary for the welfare of the cell. O2- is chemically able to reduce potentially dangerous semiquinones that might arise in the course of a cell's metabolic activities (112,113):

and interference with the detoxification of such semiquinones due to a reduction in the steady-state O2- concentration within the cell has been proposed as the basis for the harmful effects of too much superoxide dismutase. On the other hand, the Cu2+/Zn2+ dismutase is also a peroxidase, so the possibility remains that the harm caused by excess superoxide dismutase could be a result of peroxidation (114).

Regulation by O 2 - and H 2 O 2

It has become increasingly apparent over the last few years that O2- and H2O2 are signalling molecules, changing the behavior of proteins as diverse as transcription factors and membrane receptors by virtue of their ability to undergo redox reactions with the proteins with which they interact, converting -SH groups to disulfide bonds, for example, and changing the oxidation states of enzyme-associated transition metals. As signalling molecules, O2- and H2O2 are manufactured by several types of cells, including fibroblasts, endothelial and vascular smooth muscle cells, neurons, ova, spermatozoa and cells of the carotid body. All these cell types appear to use an NAD(P)H oxidase similar to the classical leukocyte NADPH oxidase to produce these oxidants. The stimuli that elicit oxidant production, however, and the purposes for which the oxidants are employed, vary from cell to cell.

1. Fibroblasts. Fibroblasts manufacture small but significant amounts of O2- in response to inflammatory mediators such as N-formylated peptides and interleukin-1 (115-117). The O2- produced by these cells has been postulated to function as a signalling molecule. Optical spectroscopy has shown that fibroblast membranes contain a heme protein that is different from the flavocytochrome subunit of the leukocyte NADPH oxidase but has properties very similar to those of the leukocyte protein (115,118). This heme protein has been suggested as the source of the O2- made by these cells.

2. Endothelial and vascular smooth muscle cells. These cells use an NAD(P)H oxidase to produce O2- in response to angiotensin II, a peptide hormone that increases blood pressure (119,120). This increase in blood pressure appears to be due to the consumption by O2- of the NO· that is generated on a continuing basis by the endothelial cells. The resulting fall in NO· concentration raises blood pressure by attenuating or eliminating the vasodilatory effect of NO· that normally prevails in the vascular tree.

3. Neurons. A recent study has shown that neuronal cells in culture produce oxidants when exposed to amyloid ß-peptide, found in amyloid deposits seen in the brains of patients with Alzheimer's disease, or related peptides from other amyloid diseases. The possibility that this O2- is produced by an NADPH oxidase is suggested by the observation that flavoprotein inhibitors known to act on the leukocyte NADPH oxidase also inhibit oxidant production in this system (121). The production of oxidants may be part of a defense used by the neuron against the peptide, with these oxidants perhaps reacting with the peptide to render it susceptible to proteolytic cleavage.

4. Ova. At the moment of fertilization, a membrane NADPH oxidase in sea urchin ova is activated to produce large amounts of H2O2 (122). This oxidant cross-links the proteins of the fertilization membrane by forming dityrosyl bridges (see Diagram 6), making the membrane impermeable to spermatozoa and thereby preventing polyspermy.

Diagram 6

5. Spermatozoa. O2- appears to be necessary for the normal function of spermatozoa. When stimulated by a calcium ionophore, normal spermatozoa generate a 3- to 5-min burst of O2- (123). The O2- produced in this reaction is involved in capacitation of the spermatozoa, because the acrosomal response to a number of stimuli is suppressed by superoxide dismutase (124). On the other hand, spermatozoa that produce O2- without stimulation are functionally abnormal, perhaps because of a generalized disruption in their signalling machinery.

6. The carotid body. The carotid body is a small organ located at the bifurcation of the common carotid artery that measures the oxygen tension of the blood (125). This organ manufactures H2O2 on a continuing basis, and immunological analysis has shown that its cells contain all 4 of the specific subunits of the leukocyte NADPH oxidase, or proteins very closely related to those subunits (126,127). It has been postulated that a carotid body NADPH oxidase very similar or identical to the leukocyte NADPH oxidase is a key component of the oxygen-measuring apparatus of the carotid body.

As to the effects of these oxidants on cellular function, there is a truly astounding number of proteins whose operation appears to depend on the redox state of the cell. Examples include the general transcription factors NF-kappa B (128-130) and AP-1 (jun/fos) (131), as well as several transcription factors that induce the synthesis of proteins that protect against oxidative stress (e.g., soxR (132,133), soxS (134), oxyR (135)). Membrane receptors and transporters, including, for example, the insulin receptor and receptors for certain neurotransmitters (136-138), are regulated by the redox state of the cell. A very large number of enzymes are also regulated by the cell's redox state. A partial list of proteins whose function is regulated by oxidation-reduction is presented in Table 1. These oxidants generally act by effecting alterations in iron-sulfur clusters (7,139-141) or by inducing the formation or rupture of disulfide bonds (142-145) on whose status the function of the protein depends. It can be postulated that at least for proteins regulated by sulfhydryl-disulfide equilibria, the effects of the oxidants are mediated through alterations in the ratio of oxidized to reduced glutathione, though this hypothesis is very difficult to prove experimentally, at least in intact cells.

Conclusion

The discovery of superoxide dismutase by McCord and Fridovich (16) has revolutionized the way biologists think about oxygen. They have come to recognize oxygen as a dangerous gift: indispensable for energy production at the level needed for living at any but the most sluggish pace, but the cause of damage that accumulates slowly over a lifetime, damage that is at least in part responsible for most of the chronic illnesses that develop with age. The challenge for the future is to develop ways to attenuate the damage inflicted by oxygen, O2- and their many descendants when they assume their evil forms.

Address for correspondence: B.M. Babior, The Scripps Research Institute, La Jolla, CA 92037, USA.

Presented at the XI Annual Meeting of the Federação de Sociedades de Biologia Experimental, Caxambu, MG, Brasil, August 21-24, 1996. Research supported in part by USPHS (Nos. AI-24227 and AI-28479). Received August 6, 1996. Accepted November 5, 1996.

  • 1. Elstner EF (1990). Der Sauerstoff. Biochemie, Biologie, Medizin BI Wissenschaftsverlag, Mannheim/Wien/Zürich.
  • 2. Fridovich I (1995). Superoxide radical and superoxide dismutases. Annual Review of Biochemistry, 64: 97-112.
  • 3. Halliwell B & Gutteridge JMC (1986). Iron and free radical reactions: two aspects of antioxidant protection. Trends in Biochemical Sciences, 11: 372-375.
  • 4. Bielski BH (1985). Fast kinetic studies of dioxygen-derived species and their metal complexes. Philosophical Transactions of the Royal Society of London, Series B. Biological Sciences, 311: 473-482.
  • 5. Goldstein S & Czapski G (1986). The role and mechanism of metal ions and their complexes in enhancing damage in biological systems or in protecting these systems from the toxicity of O2- Free Radical Biology and Medicine, 2: 3-11.
  • 6. Harris LR, Cake MH & Macey DJ (1994). Iron release from ferritin and its sensitivity to superoxide ions differs among vertebrates. Biochemical Journal, 301: 385-389.
  • 7. Gardner PR, Rainer I, Epstein LB & White CW (1995). Superoxide radical and iron modulate aconitase activity in mammalian cells. Journal of Biological Chemistry, 270: 13399-13405.
  • 8. Khan AU & Kasha M (1994). Singlet molecular oxygen in the Haber-Weiss reaction. Proceedings of the National Academy of Sciences, USA, 91: 12365-12367.
  • 9. Radi R, Beckman JS, Bush KM & Freeman BA (1991). Peroxynitrite oxidation of sulfhydryls. The cytotoxic potential of superoxide and nitric oxide. Journal of Biological Chemistry, 266: 4244-4250.
  • 10. Kong S-K, Yim MB, Stadtman ER & Chock PB (1996). Peroxynitrite disables the tyrosine phosphorylation regulatory mechanism: Lymphocyte-specific tyrosine kinase fails to phosphorylate nitrated cdc2(6-20)NH2 peptide. Proceedings of the National Academy of Sciences, USA, 93: 3377-3382.
  • 11. Winterbourn CC (1985). Comparative reactivities of various biological compounds with myeloperoxidase-hydrogen peroxide-chloride, and similarity of the oxidant to hypochlorite. Biochimica et Biophysica Acta, 840: 204-210.
  • 12. Thomas EL, Jefferson MM & Grisham M (1982). Myeloperoxidase-catalyzed incorporation of amino acids into proteins: Role of hypochlorous acid and chloramines. Biochemistry, 21: 6299-6308.
  • 13. Grisham MB, Jefferson MM, Melton DF & Thomas EL (1984). Chlorination of endogenous amines by isolated neutrophils. Ammonia-dependent bactericidal, cytotoxic, and cytolytic activities of the chloramines. Journal of Biological Chemistry, 259: 10404-10413.
  • 14. Kanofsky JR, Hoogland H, Wever R & Weiss SJ (1988). Singlet oxygen production by human eosinophils. Journal of Biological Chemistry, 263: 9692-9696.
  • 15. Steinbeck MJ, Khan AU, Karnovsky MJ & Hegg GG (1992). Intracellular singlet oxygen generation by phagocytosing neutrophils in response to particles coated with a chemical trap. Journal of Biological Chemistry, 267: 13425-13433.
  • 16. McCord JM & Fridovich I (1969). Superoxide dismutase. An enzymic function for erythrocuprein. Journal of Biological Chemistry, 244: 6049-6055.
  • 17. Halliwell B & Gutteridge JMC (1989). Free Radicals in Biology and Medicine 2nd edn. Oxford University Press, Oxford.
  • 18. Hassan HM & Fridovich I (1996). Enzymatic defenses against the toxicity of oxygen and of streptonigrin in Escherichia coli Journal of Bacteriology, 129: 1574-1583.
  • 19. Farr SB, D'Ari R & Touati D (1986). Oxygen-dependent mutagenesis in Escherichia coli lacking superoxide dismutase. Proceedings of the National Academy of Sciences, USA, 83: 8268-8272.
  • 20. Ballzan R, Bannister WH, Hunter GJ & Bannister JV (1995). Escherichia coli iron superoxide dismutase targeted to the mitochondria of yeast cells protects the cells against oxidative stress. Proceedings of the National Academy of Sciences, USA, 92: 4219-4223.
  • 21. Lapinskas PJ, Cunningham KW, Liu XF, Fink GR & Culotta VC (1995). Mutations in PMR1 suppress oxidative damage in yeast cells lacking superoxide dismutase. Molecular and Cellular Biology, 15: 1382-1388.
  • 22. Kelner MJ & Bagnell R (1990). Alteration of endogenous glutathione peroxidase, manganese superoxide dismutase, and glutathione transferase activity in cells transfected with a copper-zinc superoxide dismutase expression vector: Explanation for variations in paraquat resistance. Journal of Biological Chemistry, 265: 10872-10875.
  • 23. Yang G, Chan PH, Chen J, Carlson E, Chen SF, Weinstein P, Epstein CJ & Kamii H (1994). Human copper-zinc superoxide dismutase transgenic mice are highly resistant to reperfusion injury after focal cerebral ischemia. Stroke, 25: 165-170.
  • 24. Reveillaud I, Phillips J, Duyf B, Hilliker A, Kongpachith A & Fleming JE (1994). Phenotypic rescue by a bovine transgene in a Cu/Zn superoxide dismutase-null mutant of Drosophila melanogaster Molecular and Cellular Biology, 14: 1302-1307.
  • 25. Imlay JA & Linn S (1988). DNA damage and oxygen radical toxicity. Science, 240: 1302-1309.
  • 26. Stadtman ER (1992). Protein oxidation and aging. Science, 257: 1220-1224.
  • 27. Thomas CE, Morehouse LA & Aust SD (1985). Ferritin and superoxide-dependent lipid peroxidation. Journal of Biological Chemistry, 260: 3275-3280.
  • 28. Aikens J & Dix TA (1991). Perhydroxyl radical (HOOˇ) initiated lipid peroxidation. The role of fatty acid hydroperoxides. Journal of Biological Chemistry, 266: 15091-15098.
  • 29. Shigenaga MK, Gimeno CJ & Ames BN (1989). Urinary 8-hydroxy-2'-deoxyguanosine as a biological marker in in vivo oxidative DNA damage. Proceedings of the National Academy of Sciences, USA, 86: 9697-9701.
  • 30. Aruoma OI, Halliwell B, Gazewski E & Dizdaroglu M (1989). Damage to the bases in DNA induced by hydrogen peroxide and ferric ion chelates. Journal of Biological Chemistry, 264: 20509-20512.
  • 31. Demple B & Harrison L (1994). Repair of oxidative damage to DNA: enzymology and biology. Annual Review of Biochemistry, 63: 915-948.
  • 32. Birnboim HC & Kanabus-Kaminska M (1985). The production of DNA strand breaks in human leukocytes by superoxide anion may involve a metabolic process. Proceedings of the National Academy of Sciences, USA, 82: 6820-6824.
  • 33. Zingarelli B, O'Connor M, Wong H, Salzman AL & Szabó C (1996). Peroxynitrite-mediated DNA strand breakage activates poly-adenosine diphosphate ribosyl synthetase and causes cellular energy depletion in macrophages stimulated with bacterial lipopolysaccharide. Journal of Immunology, 156: 350-358.
  • 34. Burger RM, Projan SJ, Horwitz SB & Peisach J (1986). The DNA cleavage of iron-bleomycin. Kinetic resolution of strand scission from base propenal release. Journal of Biological Chemistry, 261: 15955-15959.
  • 35. Szabó C, Zingarelli B, O'Connor M & Salzman AL (1996). DNA strand breakage, activation of poly(ADP-ribose) synthetase, and cellular energy depletion are involved in the cytotoxicity in macrophages and smooth muscle cells exposed to peroxynitrite. Proceedings of the National Academy of Sciences, USA, 93: 1753-1758.
  • 36. Stadtman ER & Oliver CN (1991). Metal-catalyzed oxidation of proteins. Physiological consequences. Journal of Biological Chemistry, 266: 2005-2008.
  • 37. Davies KJA, Delsignore ME & Lin SW (1987). Protein damage and degradation by oxygen radicals. II. Modification of amino acids. Journal of Biological Chemistry, 262: 9902-9907.
  • 38. Stadtman ER & Berlett BS (1991). Fenton Chemistry. Amino acid oxidation. Journal of Biological Chemistry, 266: 17201-17211.
  • 39. Oliver CN, Starke-Reed PE, Stadtman ER, Liu GJ, Carney JM & Floyd RA (1990). Oxidative damage to brain proteins, loss of glutamine synthetase activity, and production of free radicals during ischemia/reperfusion-induced injury to gerbil brain. Proceedings of the National Academy of Sciences, USA, 87: 5144-5147.
  • 40. Berlett BS, Friguet B, Yim MB, Chock PB & Stadtman ER (1996). Peroxynitrite-mediated nitration of tyrosine residues in Escherichia coli glutamine synthetase mimics adenylation: Relevance to signal transduction. Proceedings of the National Academy of Sciences, USA, 93: 1776-1780.
  • 41. Haddad IY, Pataki G, Calliani C, Beckman JS & Matalon S (1994). Quantitation of nitrotyrosine levels in lung sections of patients and animals with acute lung injury. Journal of Clinical Investigation, 94: 2407-2413.
  • 42. Albrich JM, McCarthy CA & Hurst JK (1981). Biological reactivity of hypochlorous acid: Implications for microbicidal mechanisms of leukocyte myeloperoxidase. Proceedings of the National Academy of Sciences, USA, 78: 210-214.
  • 43. Domigan NM, Charlton TS, Duncan MW, Winterburn CC & Kettle AJ (1995). Chlorination of tyrosyl residues in peptides by myeloperoxidase and human neutrophils. Journal of Biological Chemistry, 270: 16542-16548.
  • 44. Bernofsky C, Bandara BMR, Hinojosa O & Strauss SL (1990). Hypochlorite-modified adenine nucleotides: Structure, spin-trapping and formation by activated guinea pig polymorphonuclear leukocytes. Free Radical Research Communications, 9: 303-315.
  • 45. Porter NA, Caldwell SE & Mills KA (1995). Mechanisms of free radical oxidation of unsaturated lipids. Lipids, 30: 277-290.
  • 46. Halliwell B (1993). The chemistry of free radicals. Toxicology and Industrial Health, 9: 1-21.
  • 47. Halliwell B & Chirico S (1993). Lipid peroxidation: Its mechanism, measurement, and significance. American Journal of Clinical Nutrition, 57: 715S-725S.
  • 48. Liu SX, Zhou M, Chen Y, Wen WY & Sun MJ (1996). Lipoperoxidative injury to macrophages by oxidatively modified low density lipoprotein may play an important role in foam cell formation. Atherosclerosis, 121: 55-61.
  • 49. Haberland ME, Fong D & Cheng L (1988). Malondialdehyde-altered protein occurs in atheroma of Watanabe heritable hyperlipidemic rabbits. Science, 241: 215-218.
  • 50. Weitzman SA & Gordon LI (1990). Inflammation and cancer: Role of phagocyte-generated oxidants in carcinogenesis. Blood, 76: 655-663.
  • 51. Floyd RA (1990). Role of oxygen free radicals in carcinogenesis and brain ischemia. FASEB Journal, 4: 2587-2597.
  • 52. Miesel R, Kurpisz M & Kroger H (1996). Suppression of inflammatory arthritis by simultaneous inhibition of nitric oxide synthase and NADPH oxidase. Free Radical Biology and Medicine, 20: 75-81.
  • 53. Adelman R, Saul RL & Ames BN (1988). Oxidative damage to DNA: Relation to species metabolic rate and life span. Proceedings of the National Academy of Sciences, USA, 85: 2706-2708.
  • 54. Ames BN, Shigenaga MK & Hagen TM (1993). Oxidants, antioxidants, and the degenerative diseases of aging. Proceedings of the National Academy of Sciences, USA, 90: 7915-7922.
  • 55. Fridovich I (1975). Superoxide dismutases. Annual Review of Biochemistry, 44: 147-159.
  • 56. Fridovich I (1974). Superoxide dismutases. Advances in Enzymology and Related Areas of Molecular Biology, 41: 35-97.
  • 57. Hosler BA & Brown Jr RH (1995). Copper/zinc superoxide dismutase mutations and free radical damage in amyotrophic lateral sclerosis. Advances in Neurology, 68: 41-46.
  • 58. Wiedau-Pazos M, Goto JJ, Rabizadeh S, Gralla EB, Roe JA, Lee MK, Valentine JS & Bredesen DE (1996). Altered reactivity of superoxide dismutase in familial amyotrophic lateral sclerosis. Science, 271: 515-518.
  • 59. Deisseroth A & Dounce AL (1970). Catalase: Physical and chemical properties, mechanism of catalysis, and physiological role. Physiological Reviews, 50: 319-375.
  • 60. Michiels C, Raes M, Toussaint O & Remacle J (1994). Importance of Se-glutathione peroxidase, catalase, and Cu/Zn-SOD for cell survival against oxidative stress. Free Radical Biology and Medicine, 17: 235-248.
  • 61. Gaetani GF, Ferraris AM, Rolfo M, Mangerini R, Arena S & Kirkman HN (1996). Predominant role of catalase in the disposal of hydrogen peroxide within human erythrocytes. Blood, 87: 1595-1599.
  • 62. Cohen HJ & Avissar N (1993). Molecular and biochemical aspects of selenium metabolism and deficiency. Progress in Clinical and Biological Research, 380: 191-202.
  • 63. Stadtman TC (1990). Selenium biochemistry. Annual Review of Biochemistry, 59: 111-127.
  • 64. Burk RF (1990). Protection against free radical injury by selenoenzymes. Pharmacology and Therapeutics, 45: 383-385.
  • 65. Flohe L (1988). Glutathione peroxidase. Basic Life Sciences, 49: 663-668.
  • 66. Chambers I & Harrison PR (1987). A new puzzle in selenoprotein biosynthesis: selenocysteine seems to be encoded by the "stop" codon, UGA. Trends in Biochemical Sciences, 12: 255-256.
  • 67. Cohen HJ, Chovaniec ME, Mistretta D & Baker SS (1985). Selenium repletion and glutathione peroxidase - differential effects on plasma and red blood cell enzyme activity. American Journal of Clinical Nutrition, 41: 735-747.
  • 68. Anonymous (1980). Treatment of glutathione peroxidase deficiency with vitamin E. Nutrition Reviews, 38: 120-122.
  • 69. Bigley R, Stankova L, Roos D & Loos J (1980). Glutathione-dependent dehydroascorbate reduction: a determinant of dehydroascorbate uptake by human polymorphonuclear leukocytes. Enzyme, 25: 200-204.
  • 70. Johnson RA, Baker SS, Fallon JT, Maynard III EP, Ruskin JN, Wen Z, Ge K & Cohen HJ (1981). An occidental case of cardiomyopathy and selenium deficiency. New England Journal of Medicine, 304: 1210-1212.
  • 71. Anonymous (1980). Prevention of Keshan cardiomyopathy by sodium selenite. Nutrition Reviews, 38: 278-279.
  • 72. Maiorino M, Chu FF, Ursini F, Davies KJA, Doroshow JH & Esworthy RS (1991). Phospholipid hydroperoxide glutathione peroxidase is the 18-kDa selenoprotein expressed in human tumor cell lines. Journal of Biological Chemistry, 266: 7728-7732.
  • 73. Frei B, England L & Ames BN (1989). Ascorbate is an outstanding antioxidant in human blood plasma. Proceedings of the National Academy of Sciences, USA, 86: 6337-6381.
  • 74. Meister A (1994). Glutathione-ascorbic acid antioxidant system in animals. Journal of Biological Chemistry, 269: 9397-9400.
  • 75. Anonymous (1989). Expanding knowledge of ascorbic acid metabolism. Nutrition Reviews, 47: 360-361.
  • 76. Levine M (1986). New concepts in the biology and biochemistry of ascorbic acid. New England Journal of Medicine, 314: 892-902.
  • 77. Chow CK (1991). Vitamin E and oxidative stress. Free Radical Biology and Medicine, 11: 215-232.
  • 78. Sies H & Murphy ME (1991). Role of tocopherols in the protection of biological systems against oxidative damage. Journal of Photochemistry and Photobiology. B, Biology, 8: 211-218.
  • 79. Packer L (1991). Protective role of vitamin E in biological systems. American Journal of Clinical Nutrition, 53: 1050S-1055S.
  • 80. Burton GW & Ingold KU (1989). Vitamin E as an in vitro and in vivo antioxidant. Annals of the New York Academy of Sciences, 570: 7-22.
  • 81. Koyama K, Takatsuki K & Inoue M (1994). Determination of superoxide and ascorbyl radicals in the circulation of animals under oxidative stress. Archives of Biochemistry and Biophysics, 309: 323-328.
  • 82. Roginsky VA & Stegmann HB (1994). Ascorbyl radical as natural indicator of oxidative stress: Quantitative regularities. Free Radical Biology and Medicine, 17: 93-103.
  • 83. Stankova L, Bigley R, Wyss SR & Aebi H (1979). Catalase and dehydroascorbate reductase in human polymorphonuclear leukocytes (PMN): possible functional relationship. Experientia, 35: 852-853.
  • 84. Mukai K, Kohno Y & Ishizu K (1988). Kinetic study of the reaction between vitamin E radical and alkyl hydroperoxides in solution. Biochemical and Biophysical Research Communications, 155: 1046-1050.
  • 85. Liebler DC, Kling DS & Reed DJ (1986). Antioxidant protection of phospholipid bilayers by alpha-tocopherol. Control of alpha-tocopherol status and lipid peroxidation by ascorbic acid and glutathione. Journal of Biological Chemistry, 261: 12114-12119.
  • 86. May JM, Qu Z & Morrow JD (1996). Interaction of ascorbate and a-tocopherol in resealed human erythrocyte ghosts. Transmembrane electron transfer and protection from lipid peroxidation. Journal of Biological Chemistry, 271: 10577-10582.
  • 87. Mukai K, Nishimura M, Ishizu K & Kitamura Y (1989). Kinetic study of the reaction of vitamin C with vitamin E radicals (tocopheroxyls) in solution. Biochimica et Biophysica Acta, 991: 276-279.
  • 88. Weiss SJ, Klein R & Slivka A (1982). Chlorination of taurine by human neutrophils. Journal of Clinical Investigation, 70: 598-607.
  • 89. Aruoma OI, Halliwell B, Hoey BM & Butler J (1988). The antioxidant action of taurine, hypotaurine and their metabolic precursors. Biochemical Journal, 256: 251-255.
  • 90. Wright CE, Lin TT, Lin YY, Sturman JA & Gaull GE (1985). Taurine scavenges oxidized chlorine in biological systems. Progress in Clinical and Biological Research, 179: 137-147.
  • 91. Weiss SJ, Klein R, Slivka A & Wei M (1982). Chlorination of taurine by human neutrophils. Evidence for hypochlorous acid generation. Journal of Clinical Investigation, 70: 598-607.
  • 92. Weiss SJ, Lampert MB & Test ST (1983). Long-lived oxidants generated by human neutrophils: Characterization and bioactivity. Science, 222: 625-628.
  • 93. Babior BM, Kipnes RS & Curnutte JT (1973). Biological defense mechanisms: the production by leukocytes of superoxide, a potential bactericidal agent. Journal of Clinical Investigation, 52: 741-744.
  • 94. Curnutte JT & Babior BM (1974). Biological defense mechanisms: the effect of bacteria and serum on superoxide production by granulocytes. Journal of Clinical Investigation, 53: 1662-1672.
  • 95. Johnston RB & Newman SL (1977). Chronic granulomatous disease. Pediatric Clinics of North America, 24: 365-376.
  • 96. Anonymous (1991). A controlled trial of interferon gamma to prevent infection in chronic granulomatous disease. The International Chronic Granulomatous Disease Cooperative Study Group. New England Journal of Medicine, 324: 509-516.
  • 97. Chanock SJ, El Benna J, Smith RM & Babior BM (1994). The respiratory burst oxidase. Journal of Biological Chemistry, 269: 24519-24522.
  • 98. Thomas EL & Fishman M (1986). Oxidation of chloride and thiocyanate by isolated leukocytes. Journal of Biological Chemistry, 261: 9694-9702.
  • 99. She Z-W, Wewers MD, Herzyk DJ, Sagone AL & Davis WB (1989). Tumor necrosis factor primes neutrophils for hypochlorous acid production. American Journal of Physiology, 257: L338-L345.
  • 100. Raschke P, Becker BF, Leipert B, Schwartz LM, Zahler S & Gerlach E (1993). Postischemic dysfunction of the heart induced by small numbers of neutrophils via formation of hypochlorous acid. Basic Research in Cardiology, 88: 321-339.
  • 101. Harrison JE & Schultz J (1976). Studies on the chlorinating activity of myeloperoxidase. Journal of Biological Chemistry, 251: 1371-1374.
  • 102. Thomas EL, Bozeman PM, Jefferson MM & King CC (1995). Oxidation of bromide by the human leukocyte enzymes myeloperoxidase and eosinophil peroxidase. Formation of bromamines. Journal of Biological Chemistry, 270: 2906-2913.
  • 103. Weiss SJ, Test ST, Eckmann CM, Roos D & Regiani S (1986). Brominating oxidants generated by human eosinophils. Science, 234: 200-202.
  • 104. Thomas EL, Grisham MB & Jefferson MM (1983). Myeloperoxidase-dependent effect of amines on functions of isolated neutrophils. Journal of Clinical Investigation, 72: 441-454.
  • 105. Rosen H, Orman J, Rakita RM, Michel BR & VanDevanter DR (1990). Loss of DNA-membrane interactions and cessation of DNA synthesis in myeloperoxidase-treated Escherichia coli Proceedings of the National Academy of Sciences, USA, 87: 10048-10052.
  • 106. Larrocha C, deCastro MF, Fontan G, Viloria A, Ferrandoz Chacon JL & Jimenez C (1982). Hereditary myeloperoxidase deficiency: a study of 12 cases. Scandinavian Journal of Haematology, 29: 389-397.
  • 107. Parry MF, Root RK, Metcalf JA, Delaney KK, Kaplow LS & Richar WJ (1981). Myeloperoxidase deficiency: Prevalence and clinical significance. Annals of Internal Medicine, 95: 293-301.
  • 108. Omar BA, Gad NM, Jordan MC, Striplin SP, Russell WJ, Downey JM & McCord JM (1990). Cardioprotection by Cu,Zn-superoxide dismutase is lost at high doses in the reoxygenated heart. Free Radical Biology and Medicine, 9: 465-471.
  • 109. Omar BA & McCord JM (1990). The cardioprotective effect of Mn-superoxide dismutase is lost at high doses in the postischemic isolated rabbit heart. Free Radical Biology and Medicine, 9: 473-478.
  • 110. Scott MD, Meshnick SR & Eaton JW (1989). Superoxide dismutase amplifies organismal sensitivity to ionizing radiation. Journal of Biological Chemistry, 264: 2498-2501.
  • 111. Scott MD, Meshnick SR & Eaton JW (1987). Superoxide dismutase-rich bacteria. Paradoxical increase in oxidant toxicity. Journal of Biological Chemistry, 262: 3640-3645.
  • 112. Winterbourn CC (1981). Cytochrome c reduction by semiquinone radicals can be indirectly inhibited by superoxide dismutase. Archives of Biochemistry and Biophysics, 209: 159-167.
  • 113. Cadenas E (1989). Biochemistry of oxygen toxicity. Annual Review of Biochemistry, 58: 79-110.
  • 114. Forage RG & Foster MA (1979). Resolution of the coenzyme B-12-dependent dehydratases of Klebsiella sp. and Citrobacter freundii Biochimica et Biophysica Acta, 569: 249-258.
  • 115. Meier B, Cross AR, Hancock JT, Kaup FJ & Jones OTG (1991). Identification of a superoxide-generating NADPH oxidase system in human fibroblasts. Biochemical Journal, 275: 241-245.
  • 116. Meier B, Radeke HH, Selle S, Habermehl GG, Resch K & Sies H (1990). Human fibroblasts release low amounts of reactive oxygen species in response to the potent phagocyte stimulants, serum-treated zymosan, N-formyl-methionyl-leucyl-phenylalanine, leukotriene B4 or 12-O-tetradecanoylphorbol 13-acetate. Biological Chemistry Hoppe-Seyler, 371: 1021-1025.
  • 117. Meier B, Radeke HH, Selle S, Younes M, Sies H, Resch K & Habermehl GG (1989). Human fibroblasts release reactive oxygen species in response to interleukin-1 or tumour necrosis factor-a. Biochemical Journal, 263: 539-545.
  • 118. Schreck R, Meier B, Mannel DN, Droge W & Baeuerle PA (1992). Dithiocarbamates as potent inhibitors of nuclear factor kappa B activation in intact cells. Journal of Experimental Medicine, 175: 1181-1194.
  • 119. Griendling KK, Minieri CA, Ollerenshaw JD & Alexander RW (1994). Angiotensin II stimulates NADH and NADPH oxidase activity in cultured vascular smooth muscle cells. Circulation Research, 74: 1141-1148.
  • 120. Rajagopalan S, Surz S, Munzel T, Tarpey M, Freeman BA, Griendling KK & Harrison DG (1996). Angiotensin II-mediated hypertension in the rat increases vascular superoxide production via membrane NADH/NADPH oxidase activation. Contribution to alterations of vasomotor tone. Journal of Clinical Investigation, 97: 1916-1923.
  • 121. Schubert D, Behl C, Lesley R, Brack A, Dargusch R, Sagara Y & Kimura H (1995). Amyloid peptides are toxic via a common oxidative mechanism. Proceedings of the National Academy of Sciences, USA, 92: 1989-1993.
  • 122. Heinecke JW & Shapiro BM (1989). Respiratory burst oxidase of fertilization. Proceedings of the National Academy of Sciences, USA, 86: 1259-1263.
  • 123. Aitken RJ & Clarkson JS (1987). Cellular basis of defective sperm function and its association with the genesis of reactive oxygen species by human spermatozoa. Journal of Reproduction and Fertility, 81: 459-469.
  • 124. De Lamirande E, Eiley D & Gagnon C (1993). Inverse relationship between the induction of human sperm capacitation and spontaneous acrosome reaction by various biological fluids and the superoxide scavenging capacity of these fluids. International Journal of Andrology, 16: 258-266.
  • 125. Acker H, Bolling B, Delpiano MA, Dufau E, Gorlach A & Holtermann G (1992). The meaning of H2O2 generation in carotid body cells for pO2 chemoreception. Journal of the Autonomic Nervous System, 41: 41-51.
  • 126. Cross AR, Henderson L, Jones OTG, Delpiano MA, Hentschel J & Acker H (1990). Involvement of an NAD(P)H oxidase as a pO2 sensor protein in the rat carotid body. Biochemical Journal, 272: 743-747.
  • 127. Kummer W & Acker H (1995). Immunohistochemical demonstration of four subunits of neutrophil NAD(P)H oxidase in type I cells of carotid body. Journal of Applied Physiology, 78: 1904-1909.
  • 128. Schreck R, Rieber P & Baeuerle PA (1991). Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-kB transcription factor and HIV-1. EMBO Journal, 10: 2247-2258.
  • 129. Menon SD, Quin S, Guy GR & Tan YH (1993). Differential induction of nuclear NF-kB by protein phosphatase inhibitors in primary and transformed human cells. Requirement for both oxidation and phosphorylation in nuclear translocation. Journal of Biological Chemistry, 268: 26805-26812.
  • 130. Baeuerle PA & Henkel T (1994). Function and activation of NF-kappaB in the immune system. Annual Review of Immunology, 12: 141-179.
  • 131. Puri PL, Avantaggiati ML, Burgio VL, Chirillo P, Collepardo D, Natoli G, Balsano C & Levrero M (1995). Reactive oxygen intermediates mediate angiotensin II-induced c-Jun.c-Fos heterodimer DNA binding activity and proliferative hypertrophic responses in myogenic cells. Journal of Biological Chemistry, 270: 22129-22134.
  • 132. Park SJ & Gunsalus RP (1995). Oxygen, iron, carbon, and superoxide control of the fumarase fumA and fumC genes of Escherichia coli: Role of the arcA, fnr, and soxR gene products. Journal of Bacteriology, 177: 6255-6262.
  • 133. Hidalgo E & Demple B (1996). Activation of SoxR-dependent transcription in vitro by noncatalytic or NifS-mediated assembly of [2Fe-2S] clusters in Apo-SoxR. Journal of Biological Chemistry, 271: 7269-7272.
  • 134. Jair KW, Fawcett WP, Fujita N, Ishihama A & Wolf Jr RE (1996). Ambidextrous transcriptional activation by SoxS: Requirement for the C-terminal domain of the RNA polymerase alpha subunit in a subset of Escherichia coli superoxide-inducible genes. Molecular Microbiology, 19: 307-317.
  • 135. Christman MF, Storz G & Ames BN (1989). OxyR, a positive regulator of hydrogen peroxide-inducible genes in Escherichia coli and Salmonella typhimurium, is homologous to a family of bacterial regulatory proteins. Proceedings of the National Academy of Sciences, USA, 86: 3484-3488.
  • 136. Marin-Hincapie M & Garofalo RS (1995). Drosophila insulin receptor: lectin-binding properties and a role for oxidative-reduction of receptor thiols in activation. Endocrinology, 136: 2357-2366.
  • 137. Pan ZH, Bahring R, Grantyn R & Lipton SA (1995). Differential modulation by sulfhydryl redox agents and glutathione of GABA- and glycine-evoked currents in rat retinal ganglion cells. Journal of Neuroscience, 15: 1384-1391.
  • 138. Staal FJT, Anderson MT, Staal GEJ, Herzenberg LA & Gitler C (1994). Redox regulation of signal transduction: Tyrosine phosphorylation and calcium influx. Proceedings of the National Academy of Sciences, USA, 91: 3619-3622.
  • 139. Hidalgo E, Bollinger Jr JM, Bradley TM, Walsh CT & Demple B (1995). Binuclear [2Fe-2S] clusters in the Escherichia coli SoxR protein and role of the metal centers in transcription. Journal of Biological Chemistry, 270: 20908-20914.
  • 140. Flint DH, Tuminello JF & Emptage MH (1993). The inactivation of Fe-S cluster containing hydrolases by superoxide. Journal of Biological Chemistry, 268: 22369-22376.
  • 141. Halliwell B (1992). Switches in enzymes. Nature, 354: 191-192.
  • 142. Bandyopadhyay S & Gronostajski RM (1994). Identification of a conserved oxidation-sensitive cysteine residue in the NF1 family of DNA-binding proteins. Journal of Biological Chemistry, 269: 29949-29955.
  • 143. Landgraf W, Regulla S, Meyer HE & Hofmann F (1991). Oxidation of cysteines activates cGMP-dependent protein kinase. Journal of Biological Chemistry, 266: 16305-16311.
  • 144. Hayashi T, Ueno Y & Okamoto T (1993). Oxidoreductive regulation of nuclear factor kappa B: Involvement of a cellular reducing catalyst thioredoxin. Journal of Biological Chemistry, 268: 11380-11388.
  • 145. Petronilli V, Constantini P, Scorrano L, Colonna R, Passamonti S & Bernardi P (1994). The voltage sensor of the mitochondrial permeability transition pore is tuned by the oxidation-reduction state of vicinal thiols. Journal of Biological Chemistry, 269: 16638-16642.
  • 146. Weiss SJ, Peppin G, Ortiz X, Ragsdale C & Test ST (1985). Oxidative autoactivation of latent collagenase by human neutrophils. Science, 227: 747-749.
  • 147. Weiss SJ & Peppin GJ (1986). Collagenolytic metalloenzymes of the human neutrophil. Characteristics, regulation and potential function in vivo. Biochemical Pharmacology, 35: 3189-3197.
  • 148. Lander HM, Ogiste JS, Teng KK & Novogrodsky A (1995). p21ras as a common signaling target of reactive free radicals and cellular redox stress. Journal of Biological Chemistry, 270: 21195-21198.
  • 149. Fialkow L, Chan CK, Grinstein S & Downey GP (1993). Regulation of tyrosine phosphorylation in neutrophils by the NADPH oxidase. Role of reactive oxygen intermediates. Journal of Biological Chemistry, 268: 17131-17137.
  • 150. Hardwick JS & Sefton MB (1995). Activation of the Lck tyrosine protein kinase by hydrogen peroxide requires the phosphorylation of Tyr-394. Proceedings of the National Academy of Sciences, USA, 92: 4527-4531.
  • 151. Ziegler DM (1985). Role of reversible oxidation-reduction of enzyme thiols-disulfides in metabolic regulation. Annual Review of Biochemistry, 54: 305-329.
  • 152. Korge P & Campbell KB (1993). The effect of changes in iron redox state on the activity of enzymes sensitive to modification of SH groups. Archives of Biochemistry and Biophysics, 304: 420-428.
  • 153. Johnson BD, Mancini-Samuelson GJ & Stankovich MT (1995). Effect of transition-state analogues on the redox properties of medium-chain acyl-CoA dehydrogenase. Biochemistry, 34: 7047-7055.
  • 154. Hassoun PM, Yu FS, Zulueta JJ, White AC & Lanzillo JJ (1995). Effect of nitric oxide and cell redox status on the regulation of endothelial cell xanthine dehydrogenase. American Journal of Physiology, 268: L809-L817.
  • 155. Li D, Stevens FJ, Schiffer M & Anderson LE (1994). Mechanism of light modulation: Identification of potential redox-sensitive cysteines distal to catalytic site in light-activated chloroplast enzymes. Biophysical Journal, 67: 29-35.
  • 156. Drincovich MF & Andreo CS (1994). Redox regulation of maize NADP-malic enzyme by thiol-disulfide interchange: effect of reduced thioredoxin on activity. Biochimica et Biophysica Acta, 1206: 10-16.
  • 157. Terada T, Nanjo H, Shinagawa K, Umemura T, Nishinaka T, Mizoguchi T & Nishihara T (1993). Modulation of 3 a-hydroxysteroid dehydrogenase activity by the redox state of glutathione. Journal of Enzyme Inhibition, 7: 33-41.
  • 158. Wunderlich M, Jaenicke R & Glockshuber R (1993). The redox properties of protein disulfide isomerase (DsbA) of Escherichia coli result from a tense conformation of its oxidized form. Journal of Molecular Biology, 233: 559-566.
  • 159. Wang GL, Jiang BH & Semenza GL (1995). Effect of altered redox states on expression and DNA-binding activity of hypoxia-inducible factor 1. Biochemical and Biophysical Research Communications, 212: 550-556.
  • 160. Arnone MI, Zannini M & Di Lauro R (1995). The DNA binding activity and the dimerization ability of the thyroid transcription factor I are redox regulated. Journal of Biological Chemistry, 270: 12048-12055.
  • 161. Esposito F, Cuccovillo F, Morra F, Russo T & Cimino F (1995). DNA binding activity of the glucocorticoid receptor is sensitive to redox changes in intact cells. Biochimica et Biophysica Acta, 1260: 308-314.
  • 162. Ammendola R, Mesuraca M, Russo T & Cimino F (1994). The DNA-binding efficiency of Sp1 is affected by redox changes. European Journal of Biochemistry, 225: 483-489.
  • 163. Gozlan H, Khazipov R & Ben-Ari Y (1995). Multiple forms of long-term potentiation and multiple regulatory sites of N-methyl-D-aspartate receptors: role of the redox site. Journal of Neurobiology, 26: 360-369.
  • 164. Sullivan JM, Traynelis SF, Chen HS, Escobar W, Heinemann SF & Lipton SA (1994). Identification of two cysteine residues that are required for redox modulation of NMDA subtype of glutamate receptor. Neuron, 13: 929-936.
  • 165. Tang LH & Aizenman E (1993). Long-lasting modification of the N-methyl-D-aspartate receptor channel by a voltage-dependent sulfhydryl redox process. Molecular Pharmacology, 44: 473-478.
  • 166. Liu G & Pessah IN (1994). Molecular interaction between ryanodine receptor and glycoprotein triadin involves redox cycling of functionally important hyperreactive sulfhydryls. Journal of Biological Chemistry, 269: 33028-33034.
  • 167. Galang CK & Hauser CA (1993). Cooperative DNA binding of the human HoxB5 (Hox-2.1) protein is under redox regulation in vitro. Molecular and Cellular Biology, 13: 4609-4617.
  • 168. Myrset AH, Bostad A, Jamin N, Lirsac PN, Toma F & Gabrielsen OS (1993). DNA and redox state induced conformational changes in the DNA-binding domain of the Myb oncoprotein. EMBO Journal, 12: 4625-4633.
  • 169. Hainaut P & Milner J (1993). Redox modulation of p53 conformation and sequence-specific DNA binding in vitro. Cancer Research, 53: 4469-4473.
  • 170. Sanchez-Garcia I & Rabbitts TH (1993). Redox regulation of in vitro DNA-binding activity by the homeodomain of the Isl-1 protein. Journal of Molecular Biology, 231: 945-949.
  • 171. Rondon IJ, Scandurro AB, Wilson RB & Beckman BS (1995). Changes in redox affect the activity of erythropoietin RNA binding protein. FEBS Letters, 359: 267-270.
  • Correspondence and Footnotes

  • Publication Dates

    • Publication in this collection
      20 Oct 1998
    • Date of issue
      Feb 1997

    History

    • Accepted
      05 Nov 1996
    • Received
      06 Aug 1996
    Associação Brasileira de Divulgação Científica Av. Bandeirantes, 3900, 14049-900 Ribeirão Preto SP Brazil, Tel. / Fax: +55 16 3315-9120 - Ribeirão Preto - SP - Brazil
    E-mail: bjournal@terra.com.br