Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 16 de 16
Filter
1.
J Biol Chem ; 292(52): 21481-21489, 2017 12 29.
Article in English | MEDLINE | ID: mdl-29109145

ABSTRACT

Cytochrome P450cam (CYP101Fe3+) regioselectively hydroxylates camphor. Possible hydroxylating intermediates in the catalytic cycle of this well-characterized enzyme have been proposed on the basis of experiments carried out at very low temperatures and shunt reactions, but their presence has not yet been validated at temperatures above 0 °C during a normal catalytic cycle. Here, we demonstrate that it is possible to mimic the natural catalytic cycle of CYP101Fe3+ by using pulse radiolysis to rapidly supply the second electron of the catalytic cycle to camphor-bound CYP101[FeO2]2+ Judging by the appearance of an absorbance maximum at 440 nm, we conclude that CYP101[FeOOH]2+ (compound 0) accumulates within 5 µs and decays rapidly to CYP101Fe3+, with a k440 nm of 9.6 × 104 s-1 All processes are complete within 40 µs at 4 °C. Importantly, no transient absorbance bands could be assigned to CYP101[FeO2+por•+] (compound 1) or CYP101[FeO2+] (compound 2). However, indirect evidence for the involvement of compound 1 was obtained from the kinetics of formation and decay of a tyrosyl radical. 5-Hydroxycamphor was formed quantitatively, and the catalytic activity of the enzyme was not impaired by exposure to radiation during the pulse radiolysis experiment. The rapid decay of compound 0 enabled calculation of the limits for the Gibbs activation energies for the conversions of compound 0 → compound 1 → compound 2 → CYP101Fe3+, yielding a ΔG‡ of 45, 39, and 39 kJ/mol, respectively. At 37 °C, the steps from compound 0 to the iron(III) state would take only 4 µs. Our kinetics studies at 4 °C complement the canonical mechanism by adding the dimension of time.


Subject(s)
Camphor 5-Monooxygenase/chemistry , Camphor 5-Monooxygenase/metabolism , Electron Transport/physiology , Binding Sites/physiology , Camphor 5-Monooxygenase/physiology , Catalysis , Cytochrome P-450 Enzyme System/metabolism , Electrons , Ferric Compounds/metabolism , Kinetics , Models, Molecular , Oxidation-Reduction
2.
Cell Physiol Biochem ; 43(4): 1533-1546, 2017.
Article in English | MEDLINE | ID: mdl-29035881

ABSTRACT

BACKGROUND/AIMS: Results from recent studies suggest that extremely low frequency magnetic fields (ELF-MF) interfere with intracellular signaling pathways related to proliferative control. The mitogen-activated protein kinases (MAPKs), central signaling components that regulate essentially all stimulated cellular processes, include the extracellular signal-regulated kinases 1/2 (ERK1/2) that are extremely sensitive to extracellular cues. Anti-phospho-ERK antibodies serve as a readout for ERK1/2 activation and are able to detect minute changes in ERK stimulation. The objective of this study was to explore whether activation of ERK1/2 and other signaling cascades can be used as a readout for responses of a variety of cell types, both transformed and non-transformed, to ELF-MF. METHODS: We applied ELF-MF at various field strengths and time periods to eight different cell types with an exposure system housed in a tissue culture incubator and followed the phosphorylation of MAPKs and Akt by western blotting. RESULTS: We found that the phosphorylation of ERK1/2 is increased in response to ELF-MF. However, the phosphorylation of ERK1/2 is likely too low to induce ELF-MF-dependent proliferation or oncogenic transformation. The p38 MAPK was very slightly phosphorylated, but JNK or Akt were not. The effect on ERK1/2 was detected for exposures to ELF-MF strengths as low as 0.15 µT and was maximal at ∼10 µT. We also show that ERK1/2 phosphorylation is blocked by the flavoprotein inhibitor diphenyleneiodonium, indicating that the response to ELF-MF may be exerted via NADP oxidase similar to the phosphorylation of ERK1/2 in response to microwave radiation. CONCLUSIONS: Our results further indicate that cells are responsive to ELF-MF at field strengths much lower than previously suspected and that the effect may be mediated by NADP oxidase. However, the small increase in ERK1/2 phosphorylation is probably insufficient to affect proliferation and oncogenic transformation. Therefore, the results cannot be regarded as proof of the involvement of ELF-MF in cancer in general or childhood leukemia in particular.


Subject(s)
Electromagnetic Fields , Enzyme Activation , MAP Kinase Signaling System , Animals , Cell Line , Cell Line, Tumor , Electromagnetic Fields/adverse effects , Humans , Mitogen-Activated Protein Kinase 1/metabolism , Mitogen-Activated Protein Kinase 3/metabolism , NADPH Oxidases/metabolism , Neoplasms/etiology , Neoplasms/metabolism , Phosphorylation
3.
Arch Biochem Biophys ; 617: 3-8, 2017 Mar 01.
Article in English | MEDLINE | ID: mdl-27670814

ABSTRACT

There is currently interest in sulfur-containing molecules that may or may not play a role in signaling. We have collected relevant thermodynamic data, namely standard Gibbs energies of formation and electrode potentials at pH 7, and used these to construct a Frost diagram. Thermodynamic data not available in the literature could be estimated with reasonable confidence. At pH 7, the electrode potential of the RSS/RSS- couple is +0.68 V, 0.28 V less than that of the RS, H+/RSH couple. S2- is unstable with respect to HSS- and S2. Generally, polysulfur compounds, with the exception of RSSR, are thermodynamically unstable with respect to disproportionation and ultimately lead to formation of kinetically inert S8(s). About thermoneutral is the formation of RSS- from RSSR and HS-, but formation of HNO from HS- and SNO-, and from HS- and RSNO, is unfavorable. The formation of SSNO-in vivo is kinetically unlikely.


Subject(s)
Biochemistry/methods , Signal Transduction , Sulfur/chemistry , Electrodes , Electrons , Hydrogen-Ion Concentration , Kinetics , Oxidation-Reduction , Oxidative Stress , Oxygen/chemistry , Sulfhydryl Compounds/chemistry , Temperature , Thermodynamics
4.
Bioelectromagnetics ; 37(3): 183-189, 2016 Apr.
Article in English | MEDLINE | ID: mdl-26991812

ABSTRACT

Exposure to extremely low-frequency magnetic fields (ELF-MF) was evaluated in an International Agency for Research on Cancer (IARC) Monographs as "possibly carcinogenic to humans" in 2001, based on increased childhood leukemia risk observed in epidemiological studies. We conducted a hazard assessment using available scientific evidence published before March 2015, with inclusion of new research findings from the Advanced Research on Interaction Mechanisms of electroMagnetic exposures with Organisms for Risk Assessment (ARIMMORA) project. The IARC Monograph evaluation scheme was applied to hazard identification. In ARIMMORA for the first time, a transgenic mouse model was used to mimic the most common childhood leukemia: new pathogenic mechanisms were indicated, but more data are needed to draw definitive conclusions. Although experiments in different animal strains showed exposure-related decreases of CD8+ T-cells, a role in carcinogenesis must be further established. No direct damage of DNA by exposure was observed. Overall in the literature, there is limited evidence of carcinogenicity in humans and inadequate evidence of carcinogenicity in experimental animals, with only weak supporting evidence from mechanistic studies. New exposure data from ARIMMORA confirmed that if the association is nevertheless causal, up to 2% of childhood leukemias in Europe, as previously estimated, may be attributable to ELF-MF. In summary, ARIMMORA concludes that the relationship between ELF-MF and childhood leukemia remains consistent with possible carcinogenicity in humans. While this scientific uncertainty is dissatisfactory for science and public health, new mechanistic insight from ARIMMORA experiments points to future research that could provide a step-change in future assessments. Bioelectromagnetics. 37:183-189, 2016. © 2016 Wiley Periodicals, Inc.

5.
Biochemistry ; 54(3): 711-21, 2015 Jan 27.
Article in English | MEDLINE | ID: mdl-25545162

ABSTRACT

Cytochrome P450 3A4 (CYP3A4) is the major human P450 responsible for the metabolism of carbamazepine (CBZ). To explore the mechanisms of interactions of CYP3A4 with this anticonvulsive drug, we carried out multiple molecular dynamics (MD) simulations, starting with the complex of CYP3A4 manually docked with CBZ. On the basis of these simulations, we engineered CYP3A4 mutants I369F, I369L, A370V, and A370L, in which the productive binding orientation was expected to be stabilized, thus leading to increased turnover of CBZ to the 10,11-epoxide product. In addition, we generated CYP3A4 mutant S119A as a control construct with putative destabilization of the productive binding pose. Evaluation of the kinetics profiles of CBZ epoxidation demonstrate that CYP3A4-containing bacterial membranes (bactosomes) as well as purified CYP3A4 (wild-type and mutants I369L/F) exhibit substrate inhibition in reconstituted systems. In contrast, mutants S119A and A370V/L exhibit S-shaped profiles that are indicative of homotropic cooperativity. MD simulations with two to four CBZ molecules provide evidence that the substrate-binding pocket of CYP3A4 can accommodate more than one molecule of CBZ. Analysis of the kinetics profiles of CBZ metabolism with a model that combines the formalism of the Hill equation with an allowance for substrate inhibition demonstrates that the mechanism of interactions of CBZ with CYP3A4 involves multiple substrate-binding events (most likely three). Despite the retention of the multisite binding mechanism in the mutants, functional manifestations reveal an exquisite sensitivity to even minor structural changes in the binding pocket that are introduced by conservative substitutions such as I369F, I369L, and A370V.


Subject(s)
Carbamazepine/metabolism , Catalytic Domain , Cytochrome P-450 CYP3A/metabolism , Epoxy Compounds/metabolism , Molecular Dynamics Simulation , Mutant Proteins/metabolism , Carbamazepine/chemistry , Cytochrome P-450 CYP3A/chemistry , Cytochrome P-450 CYP3A/isolation & purification , Heme/metabolism , Humans , Mutant Proteins/chemistry , Mutation , Substrate Specificity
6.
Chem Res Toxicol ; 28(4): 604-14, 2015 Apr 20.
Article in English | MEDLINE | ID: mdl-25654270

ABSTRACT

Iron in iron overload disease is present as non-transferrin-bound iron, consisting of iron, citrate, and albumin. We investigated the redox properties of iron citrate by electrochemistry, by the kinetics of its reaction with ascorbate, by ESR, and by analyzing the products of reactions of ascorbate with iron citrate complexes in the presence of H2O2 with 4-hydroxybenzoic acid as a reporter molecule for hydroxylation. We report -0.03 V < E°' > +0.01 V for the (Fe(3+)-cit/Fe(2+)-cit) couple. The first step in the reaction of iron citrate with ascorbate is the rapid formation of mixed complexes of iron with citrate and ascorbate, followed by slow reduction to Fe(2+)-citrate with k = ca. 3 M(-1) s(-1). The ascorbyl radical is formed by iron citrate oxidation of Hasc(-) with k = ca. 0.02 M(-1) s(-1); the majority of the ascorbyl radical formed is sequestered by complexation with iron and remains EPR silent. The hydroxylation of 4-hydroxybenzoic acid driven by the Fenton reduction of iron citrate by ascorbate in the presence of H2O2 proceeds in three phases: the first phase, which is independent of the presence of O2, is revealed as a nonzero intercept that reflects the rapid reaction of accumulated Fe(2+) with H2O2; the intermediate oxygen-dependent phase fits a first-order accumulation of product with k = 5 M(-1) s(-1) under aerobic and k = 13 M(-1) s(-1) under anaerobic conditions; the slope of the final linear phase is ca. k = 5 × 10(-2) M(-1) s(-1) under both aerobic and anaerobic conditions. Product yields under aerobic conditions are greater than predicted from the initial concentration of iron, but they are less than predicted for continuous redox cycling in the presence of excess ascorbate. The ongoing formation of hydroxylated product supports slow redox cycling by iron citrate. Thus, when H2O2 is available, iron-citrate complexes may contribute to pathophysiological manifestations of iron overload diseases.


Subject(s)
Citric Acid/chemistry , Iron/chemistry , Electrochemical Techniques , Electron Spin Resonance Spectroscopy , Kinetics , Oxidation-Reduction
8.
Amino Acids ; 39(5): 1131-7, 2010 Nov.
Article in English | MEDLINE | ID: mdl-20532951

ABSTRACT

The oxidation of proteins and other macromolecules by radical species under conditions of oxidative stress can be modulated by antioxidant compounds. Decreased levels of the antioxidants glutathione and ascorbate have been documented in oxidative stress-related diseases. A radical generated on the surface of a protein can: (1) be immediately and fully repaired by direct reaction with an antioxidant; (2) react with dioxygen to form the corresponding peroxyl radical; or (3) undergo intramolecular long range electron transfer to relocate the free electron to another amino acid residue. In pulse radiolysis studies, in vitro production of the initial radical on a protein is conveniently made at a tryptophan residue, and electron transfer often leads ultimately to residence of the unpaired electron on a tyrosine residue. We review here the kinetics data for reactions of the antioxidants glutathione, selenocysteine, and ascorbate with tryptophanyl and tyrosyl radicals as free amino acids in model compounds and proteins. Glutathione repairs a tryptophanyl radical in lysozyme with a rate constant of (1.05±0.05)×10(5) M(-1) s(-1), while ascorbate repairs tryptophanyl and tyrosyl radicals ca. 3 orders of magnitude faster. The in vitro reaction of glutathione with these radicals is too slow to prevent formation of peroxyl radicals, which become reduced by glutathione to hydroperoxides; the resulting glutathione thiyl radical is capable of further radical generation by hydrogen abstraction. Although physiologically not significant, selenoglutathione reduces tyrosyl radicals as fast as ascorbate. The reaction of protein radicals formed on insulin, ß-lactoglobulin, pepsin, chymotrypsin and bovine serum albumin with ascorbate is relatively rapid, competes with the reaction with dioxygen, and the relatively innocuous ascorbyl radical is formed. On the basis of these kinetics data, we suggest that reductive repair of protein radicals may contribute to the well-documented depletion of ascorbate in living organisms subjected to oxidative stress.


Subject(s)
Amino Acids/metabolism , Ascorbic Acid/metabolism , Glutathione/metabolism , Proteins/metabolism , Free Radicals/metabolism , Oxidation-Reduction
9.
Phys Chem Chem Phys ; 12(37): 11609-16, 2010 Oct 07.
Article in English | MEDLINE | ID: mdl-20689867

ABSTRACT

An understanding of the reactivity of oligomeric compounds that model fuel cell membrane materials under oxidative-stress conditions that mimic the fuel cell operating environment can identify material weaknesses and yield valuable insights into how a polymer might be modified to improve oxidative stability. The reaction of HO˙ radicals with a polymer electrolyte fuel cell membrane represents an initiation step for irreversible membrane oxidation. By means of pulse radiolysis, we measured k = (9.5 ± 0.6) × 10(9) M(-1) s(-1) for the reaction of HO˙ with poly(sodium styrene sulfonate), PSSS, with an average molecular weight of 1100 Da (PSSS-1100) in aqueous solution at room temperature. In the initial reaction of HO˙ with the oligomer (90 ± 10)% react by addition to form hydroxycyclohexadienyl radicals, while the remaining abstract a hydrogen to yield benzyl radicals. The hydroxycyclohexadienyl radicals react reversibly with dioxygen to form the corresponding peroxyl radicals; the second-order rate constant for the forward reaction is k(f) = (3.0 ± 0.5) × 10(7) M(-1) s(-1), and for the back reaction, we derive an upper limit for the rate constant k(r) of (4.5 ± 0.9) × 10(3) s(-1). These data place a lower bound on the equilibrium constant K of (7 ± 2) × 10(3) M(-1) at 295 K, which allows us to calculate a lower limit of the Gibbs energy for the reaction, (-21.7 ± 0.8) kJ mol(-1). At pH 1, the hydroxycyclohexadienyl radicals decay with an overall first-order rate constant k of (6 ± 1) × 10(3) s(-1) to yield benzyl radicals. The second-order rate constant for reaction of dioxygen with benzyl radicals of PSSS-1100 is k = (2-5) × 10(8) M(-1) s(-1). We discuss hydrogen abstraction from PSSS-1100 in terms of the bond dissociation energy, and relate these to relevant electrode potentials. We propose a reaction mechanism for the decay of hydroxycyclohexadienyl radicals and subsequent reaction steps.

10.
Biochemistry ; 47(36): 9602-7, 2008 Sep 09.
Article in English | MEDLINE | ID: mdl-18702524

ABSTRACT

The rate constant for the reduction of the tyrosyl radical with selenocysteine has been measured to investigate whether selenocysteine is capable of repair of protein radicals. Tyrosyl radicals, both free in solution and in insulin, were generated by means of pulse radiolysis and laser flash photolysis in aqueous solution. The rate constant for the reaction of free N-acetyl-tyrosyl-amine radicals with selenocysteine is (8 +/- 2) x 10 (8) M (-1) s (-1), and that for tyrosyl radicals in insulin is (1.6 +/- 0.4) x 10 (8) M (-1) s (-1). The rate constant for the reaction of selenoglutathione with the N-acetyl-tyrosyl-amine radical is (5 +/- 2) x 10 (8) M (-1) s (-1). In contrast, cysteine and glutathione react more slowly than their selenium analogues with the tyrosyl radical: the reactions of N-acetyl-tyrosyl-amine radicals with cysteine and glutathione are 3 and 5 orders of magnitude slower, respectively, than those with selenocysteine and selenoglutathione, while those of tyrosyl radicals in insulin are 3 and 2 orders of magnitude slower, respectively.


Subject(s)
Free Radicals/chemistry , Selenocysteine/chemistry , Tyrosine/chemistry , Photolysis , Pulse Radiolysis/methods
11.
Dalton Trans ; 41(45): 13779-87, 2012 Dec 07.
Article in English | MEDLINE | ID: mdl-23007920

ABSTRACT

The isomerisation of ONOOH to NO(3)(-) and H(+), some oxidations and all hydroxylations and nitrations of aromatic compounds are first-order in ONOOH and zero-order in the compounds that are modified. These reactions are widely believed to proceed via homolysis of ONOOH into HO˙ and NO(2)˙ to an extent of ca. 30%. We review the evidence pro and contra homolysis in studies that involve (1) thermochemical considerations, (2) isomerisation to NO(3)(-) and H(+), (3) decomposition to NO(2)(-) and O(2), (4) HO˙ scavenger studies, (5) deuterium isotope effects, (6) (18)O-scrambling studies, (7) electrochemistry, (8) CIDNP NMR, and (9) photolysis. Our conclusion is that homolysis may be involved to a minor extent of ca. 5%. The initiation of ONOOH isomerisation may be visualised as an out-of-plane vibration of the terminal HO-group relative to the nitrogen. At ONOO(-) concentrations exceeding 0.1 mM and near neutral pH, disproportionation to NO(2)(-) and O(2) occurs; such disproportionations are typical for peroxy acids. For oxidation and nitration of organic substrates, we favour a mechanism involving initial formation of an adduct between the compound to be oxidised or nitrated and ONOOH.


Subject(s)
Oxidants/chemistry , Peroxynitrous Acid/chemistry , Molecular Structure , Stereoisomerism , Temperature
12.
Nat Rev Cancer ; 11(5): 325-37, 2011 May.
Article in English | MEDLINE | ID: mdl-21508971

ABSTRACT

Otto Warburg pioneered quantitative investigations of cancer cell metabolism, as well as photosynthesis and respiration. Warburg and co-workers showed in the 1920s that, under aerobic conditions, tumour tissues metabolize approximately tenfold more glucose to lactate in a given time than normal tissues, a phenomenon known as the Warburg effect. However, this increase in aerobic glycolysis in cancer cells is often erroneously thought to occur instead of mitochondrial respiration and has been misinterpreted as evidence for damage to respiration instead of damage to the regulation of glycolysis. In fact, many cancers exhibit the Warburg effect while retaining mitochondrial respiration. We re-examine Warburg's observations in relation to the current concepts of cancer metabolism as being intimately linked to alterations of mitochondrial DNA, oncogenes and tumour suppressors, and thus readily exploitable for cancer therapy.


Subject(s)
Neoplasms/history , Cell Respiration/physiology , Germany , Glycolysis/physiology , History, 20th Century , Humans , Neoplasms/metabolism , Neoplasms/pathology
13.
Radiat Res ; 176(1): 128-33, 2011 Jul.
Article in English | MEDLINE | ID: mdl-21631291

ABSTRACT

3-Nitrotyrosine has been reported as an important biomarker of oxidative stress that may play a role in a variety of diseases. In this work, transient UV-visible absorption spectra and kinetics observed during the reaction of the hydrated electron, e(aq)(-), with 3-nitrotyrosine and derivatives thereof were investigated. The absorption spectra show characteristics of aromatic nitro anion radicals. The absorptivity of radical anion product at 300 nm is estimated to be (1.0 ± 0.2) × 10(4) M(-1) cm(-1) at pH 7.3. The rate constants determined for the reaction of e(aq)(-) with 3-nitrotyrosine, N-acetyl-3-nitrotyrosine ethyl ester and glycylnitrotyrosylglycine at neutral pH (3.0 ± 0.3) × 10(10) M(-1) s(-1), (2.9 ± 0.2) × 10(10) M(-1) s(-1) and (1.9 ± 0.2) × 10(10) M(-1) s(-1), respectively, approach the diffusion-control limit and are almost two orders of magnitude higher than those for the reactions with tyrosine and tyrosine-containing peptides. The magnitude of the rate constants supports reaction of e(aq)(-) at the nitro group, and the product absorbance at 300 nm is consistent with formation of the nitro anion radical. The pH dependence of the second-order rate constant for e(aq)(-) decay (720 nm) in the presence of 3-nitrotyrosine shows a decrease with increasing pH, consistent with unfavorable electrostatic interactions. The pH dependence of the second-order rate constant for formation of radical anion (300 nm) product suggests that deprotonation of the amino group slows the rate, which indicates that deamination to form the 1-carboxy-2-(4-hydroxy-3-nitrophenyl)ethyl radical occurs. We conclude that the presence of the nitro group activates tyrosine and derivatives toward reaction with e(aq)(-) and can affect the redox chemistry of biomolecules exposed to oxidative stress.


Subject(s)
Electrons , Tyrosine/analogs & derivatives , Absorption , Kinetics , Oligopeptides/chemistry , Oligopeptides/metabolism , Pulse Radiolysis , Spectrum Analysis , Tyrosine/metabolism
14.
Free Radic Biol Med ; 49(3): 317-22, 2010 Aug 01.
Article in English | MEDLINE | ID: mdl-20406682

ABSTRACT

We use published Gibbs energies of formation and equilibrium constants to determine electrode potentials for the partially reduced intermediates along the pathway of reduction of dioxygen to water, as well as of ozone and singlet dioxygen. The results are summarized in an oxidation state (Frost) diagram. Our review of the literature on electrode potentials leads us to revise values for the O(2)/O(2)(*-) couple to E degrees (O(2g)/O(2)(*-))=-0.35+/-0.02V and E degrees (O(2aq)/O(2)(*-))=-0.18+/-0.02V from -0.33 and -0.16V, respectively. Other electrode potentials (pH 7) for the radical species covered are E degrees '(O(3g)/O(3)(*-))=+0.91V, E degrees '(HO(2)(*), H(+)/H(2)O(2))=+1.05V, E degrees '(H(2)O(2), H(+)/HO(*), H(2)O)=+0.39V, and E degrees '(HO(*), H(+)/H(2)O)=+2.31V.


Subject(s)
Oxygen/chemistry , Electrodes , Hydrogen Peroxide/chemistry , Hydroxides/chemistry , Oxidation-Reduction , Oxygen/metabolism , Thermodynamics , Water/chemistry
16.
Arch Biochem Biophys ; 439(2): 234-40, 2005 Jul 15.
Article in English | MEDLINE | ID: mdl-15978540

ABSTRACT

The kinetics of bovine Cu,Zn superoxide dismutase were studied by pulse radiolysis. To ensure the absence of catalytically active free copper, commercially obtained holo-superoxide dismutase was demetallated, and the apo-superoxide dismutase concentrations were determined by isothermal titration calorimetry prior to reconstitution with defined amounts of copper and zinc. The catalytic rate constant was determined as a function of ionic strength over the range of 4-154 mM, and of the copper and zinc content. The catalytic rate constant increases with ionic strength up to (1.5 +/- 0.2) x 10(9) M(-1) s(-1) at an ionic strength of 15 mM, and then decreases. At pH 7 and 50 mM ionic strength, k = (1.2 +/- 0.2) x 10(9) M(-1) s(-1), and at a physiologically relevant ionic strength of 150 mM, it is (0.7 +/- 0.1) x 10 (9) M(-1) s(-1). The effect of ionic strength is ascribed to the inhomogeneous electric field generated by the surface charges of superoxide dismutase. The value of the catalytic rate constant at 50 mM is ca. 2-fold smaller than earlier values reported in the literature. The relationship between copper content and the catalytic rate constant shows that addition of more than a stoichiometric amount of copper cannot be masked efficiently by EDTA. The possibility exists that earlier reported values were based on experiments contaminated with trace amounts of copper.


Subject(s)
Copper/metabolism , Superoxide Dismutase/metabolism , Zinc/metabolism , Animals , Cattle , Copper/chemistry , Edetic Acid/chemistry , Kinetics , Osmolar Concentration , Pulse Radiolysis , Superoxide Dismutase/chemistry , Zinc/chemistry
SELECTION OF CITATIONS
SEARCH DETAIL