首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The reactions of RO(2)* radicals with Fe(H(2)O)(6)(2+) were studied, R[double bond]H; CH(3); CH(2)COOH; CH(2)CN; CH(2)C(CH(3))(2)OH; CH(2)OH; CHCl(2)/CCl(3). All these processes involve the following reactions: Fe(H(2)O)(6)(2+)+RO(2)*<==>(H(2)O)(5)Fe(III)[bond]OOR(2+) K(1) approximately 250 M(-1); (H(2)O)(5)Fe(III)[bond]OOR(2+)+H(3)O(+)/H(2)O-->Fe(H(2)O)(6)(3+)+ROOH+H(2)O/OH(-); (H(2)O)(5)Fe(III)[bond]OOR(2+)+2Fe(H(2)O)(6)(2+)-->3Fe(H(2)O)(6)(3+)+ROH; 2 RO(2)*-->Products; RO(2)*+(H(2)O)(5)Fe(III)[bond]OOR(2+)-->Fe(H(2)O)(6)(2+)+products. The values of k(1) and k(3) [reaction is clearly not an elementary reaction] approach the ligand exchange rate of Fe(H(2)O)(6)(2+), i.e. these reactions follow an inner sphere mechanism and the rate determining step is the ligand exchange step. The rate of reaction is several orders of magnitude faster than that of the Fenton reaction. Surprisingly enough the K(1) values are nearly independent of the redox potential of the radical and are considerably higher than calculated from the relevant redox potentials. These results indicate that the ROO(-) ligands considerably stabilise the Fe(III) complex, this stabilisation is smaller for radicals with electron withdrawing groups which raise the redox potential of the radical but decrease the basicity of the ROO(-) ligands, two effects which seem to nearly cancel each other. Finally, the results clearly indicate that reaction (5) is relatively fast and affects the nature of the final products. The contribution of these reactions to oxidation processes involving 'Fenton-like' processes is discussed.  相似文献   

2.
Three new binary Cu(II) complexes of norfloxacin have been synthesized and characterized. We also report the synthesis, characterization and X-ray crystallographic structures of a new binary compound, [Cu(HNor)(2)]Cl(2).2H(2)O (2) and two new ternary complexes norfloxacin-copper(II)-phen, [Cu(Nor)(phen)(H(2)O)](NO(3)).3H(2)O (4), and [Cu(HNor)(phen)(NO(3))](NO(3)).3H(2)O (5). The structure of 2 consists of two crystallographically independent cationic monomeric units of [Cu(HNor)(2)](2+), chloride anions, and uncoordinated water molecules. The Cu(II) ion is placed at a center of symmetry and is coordinated to two norfloxacin ligands which are related through the inversion center. The structures of 4 and 5 consist of cationic units ([Cu(Nor)(phen)(H(2)O)](+) for 4 and [Cu(HNor)(phen)(NO(3))](+) for 5), nitrate counteranions, and lattice water molecules that provide crystalline stability through a network of hydrogen-bond interactions. The complexes exhibit a five coordinated motif in a square pyramidal environment around the metal center. The ability of compounds 4 and 5 to cleave DNA has also been studied. Mechanistic studies with different inhibiting reagents reveal that hydroxyl radicals, singlet oxygen, and superoxide radicals are all involved in the DNA scission process mediated by these compounds.  相似文献   

3.
The cytotoxicity of certain Cr(III) complexes, such as [Cr(salen)(H(2)O)(2)](+), [Cr(edta)(H(2)O)](-), [Cr(en)(3)](3+), [Cr(ox)(3)](3-), [Cr(pic)(3)], and CrCl(3), which differ in ionic character and ligand environment in human dermal skin fibroblasts, has been studied. After 72 h of exposure to 100 microM doses of chromium(III) complexes, the order in which the complexes had an inhibitory effect on cell viability was [Cr(en)(3)](3+) > [Cr(salen)(H(2)O)(2)](+) > [Cr(ox)(3)](3-) > [Cr(edta)(H(2)O)](-) > [Cr(pic)(3)] > CrCl(3). Based on viability studies it was confirmed that [Cr(en)(3)](3+), a triply charged cation, inhibits cell proliferation, and therefore, it was chosen to carry out further investigations. [Cr(en)(3)](3+), at a dose of 50 microM, was found to bring about surface morphological changes, evidenced by cellular blebbing and spike formation accompanied by nuclear damage. TEM analysis revealed substantial intracellular damage to fibroblasts in terms of the formation of apoptotic bodies and chromatin condensation, thus reflecting cell death. FACS analysis further revealed DNA damage by formation of a sub-G(1) peak with 84.2% DNA as aneuploid DNA and arrest of the G(2) / M phase of the cell cycle. Cellular DNA damage was confirmed by agarose gel electrophoresis with the characteristic appearance of a DNA streak in DNA isolated from [Cr(en)(3)](3+)-treated fibroblasts. The proposed mechanism suggests the plausible role of Cr(V), formed as a result of oxidation of Cr(III) by cellular oxidative enzymes, in the cytotoxic response. Consequently, any Cr(III) complex that is absorbed by cells and can be oxidized to Cr(V) must be considered a potential carcinogen. This has potential implications for the increased use of Cr(III) complexes as dietary supplements and highlights the need to consider the cytotoxicity and genotoxicity of a variety of Cr(III) complexes and to understand the potential hazards of Cr(III) complexes encountered in research laboratories.  相似文献   

4.
Y Liu  K Zhang  Y Wu  J Zhao  J Liu 《Chemistry & biodiversity》2012,9(8):1533-1544
8-Hydroxyquinoline-7-carboxaldehyde (8-HQ-7-CA), Schiff-base ligand 8-hydroxyquinoline-7-carboxaldehyde benzoylhydrazone, and binuclear complexes [LnL(NO(3) )(H(2) O)(2) ](2) were prepared from the ligand and equivalent molar amounts of Ln(NO(3) )?6 H(2) O (Ln=La(3+) , Nd(3+) , Sm(3+) , Eu(3+) , Gd(3+) , Dy(3+) , Ho(3+) , Er(3+) , Yb(3+) , resp.). Ligand acts as dibasic tetradentates, binding to Ln(III) through the phenolate O-atom, N-atom of quinolinato unit, and C?N and ?O?C?N? groups of the benzoylhydrazine side chain. Dimerization of this monomeric unit occurs through the phenolate O-atoms leading to a central four-membered (LnO)(2) ring. Ligand and all of the Ln(III) complexes can strongly bind to CT-DNA through intercalation with the binding constants at 10(5) -10(6) M(-1) . Moreover, ligand and all of the Ln(III) complexes have strong abilities of scavenging effects for hydroxyl (HO(.) ) radicals. Both the antioxidation and DNA-binding properties of Ln(III) complexes are much better than that of ligand.  相似文献   

5.
We examined whether superoxide (O(2)(-)) is produced as a precursor of hydrogen peroxide (H(2)O(2)) in cultured thyroid cells using the cytochrome c method and the electron paramagnetic resonance (EPR) method. No O(2)(-) or its related radicals was detected in thyroid cells under the physiological condition. The presence of quinone, 2,3-dimethoxy-l-naphthoquinone (DMNQ), or 2-methyl-1, 4-naphthoquinone (menadione), in the medium produced O(2)(-) and hydroxyl radicals (OH*); the amount of H(2)O(2) generation was also increased. Incubation of follicles with DMNQ or menadione inhibited iodine organification (a step of thyroid hormone formation) and its catalytic enzyme, thyroid peroxidase (TPO). This inhibition should be caused by reactive oxygen species because the two quinones, particularly DMNQ, exert their effect through the generation of reactive oxygen species. It is speculated that the site-specific inactivation of TPO might have occurred at the heme-linked histidine residue of the TPO molecule, a critical amino acid for enzyme activity because OH* (vicious free radicals) can be formed at the iron-linked amino acid. TPO mRNA level and electrophoretic mobility of TPO were not inhibited by quinones. Our study suggests that thyroid H(2)O(2) is produced by divalent reduction of oxygen without O(2)(-) generation. If thyroid cells happen to be exposed to significant amount of reactive oxygen species, TPO and subsequent thyroid hormone formation are inhibited.  相似文献   

6.
Formation of reduced nicotinamide adenine dinucleotide peroxide   总被引:1,自引:0,他引:1  
Incubation of NADH at neutral and slightly alkaline pH leads to the gradual absorption of 1 mol of H+. This uptake of acid requires oxygen and mainly yields anomerized NAD+ (NAD+), with only minimal formation od acid-modified NADH. The overall stoichiometry of the reaction is: NADH + H+ + 1/2O2 leads to H2O + NAD+, with NADH peroxide (HO2-NADH+) serving as the intermediate that anomerizes and breaks down to give NAD+ and H2O2. The final reaction reaction mixture contains less than 0.1% of the generated H2O2, which is nonenzymically reduced by NADH. The latter reaction is inhibited by catalase, leading to a decrease in the overall rate of acid absorption, and stimulated by peroxidase, leading to an increase in the overall rate of acid absorption. Although oxygen can attack NADH at either N-1 or C-5 of the dihydropyridine ring, the attack appears to occur primarily at N-1. This assignment is based on the inability of the C-5 peroxide to anomerize, whereas the N-1 peroxide, being a quaternary pyridinium compound, can anomerize via reversible dissociation of H2O2. The peroxidase-catalyzed oxidation of NADH by H2O2 does not lead to anomerization, indicating that anomerization occurs prior to the release of H2O2. Chromatography of reaction mixtures on Dowex 1 formate shows the presence of two major and several minor neutral and cationic degradation products. One of the major products is nicotinamide, which possibly arises from breakdown of nicotinamide-1-peroxide. The other products have not been identified, but may be derived from other isomeric nicotinamide peroxides.  相似文献   

7.
Oxidative DNA damage was investigated by free radicals generated from HEPES (2-[4-(2-hydroxyethyl)-1-piperazinyl]ethanesulfonic acid) buffer, which is widely used in biochemical or biological studies, in the presence of Au(III). The effect of free radicals on the DNA damage was ascertained by gel electrophoresis, electron spin resonance (ESR) spectroscopy and circular dichroism (CD) spectroscopy. ESR results indicated the generation of nitrogen-centered cationic free radicals from the HEPES in the presence of Au(III) which cause the DNA damage. No ESR spectra were observed for phosphate, tris(hydroxymethyl)aminomethane (Tris-HCl) and acetate buffers in the presence of Au(III) or for HEPES buffer in the presence of other metal ions such as Mn(II), Fe(III), Co(II), Ni(II), Cu(II), Zn(II) and Pd(II) or [Au(III)(TMPyP)](5+) and [Pd(II)(TMPyP)](4+), where [H(2)(TMPyP)](4+) denotes tetrakis(1-methylpyridium-4-yl)porphyrin. Consequently, no DNA damage was observed for these buffer agents (e.g., phosphate, Tris-HCl or acetate) in the presence of Au(III) or for HEPES in the presence of other metal ions or the metalloporphyrins mentioned above. No detectable inhibitory effect on the DNA damage was observed by using the typical scavengers of reactive oxygen species (ROS) ()OH, O(2)(-) and H(2)O(2). This non-inhibitory effect indicated that no reactive oxygen species were generated during the incubation of DNA with HEPES and Au(III). The drastic change in CD spectra from positive ellipticity to negative ellipticity approximately at 270 nm with increasing concentration of Au(III) also indicated the significant damage of DNA. Only HEPES or Au(III) itself did not damage DNA. A mechanism for the damaging of DNA is proposed.  相似文献   

8.
The effect of low concentrations of sodium oleate on the oxidation of oxymyoglobin to metmyoglobin has been examined. This long chain fatty acid results in a tripling of the initial rate (1.5-4.3 h-1) at which oxymyoglobin is converted to metmyoglobin and more than doubling of the rate of the long-term reaction (0.12-0.33 h-1). Examination of rate constant enhancement over a range of oleate concentrations (0-0.215 mM) has allowed an estimate of association constants for both phases of the reaction system. The peroxidase activity expressed by metmyoglobin towards hydrogen peroxide is inhibited by the presence of sodium oleate by a fivefold increase in the apparent Km value (0.33-1.77 mM). The observed changes in oxymyoglobin concentration over time are discussed in terms of competition between metmyoglobin, which acts as a peroxidase decreasing in situ concentrations of H2O2, and oxymyoglobin, which also is oxidized by the peroxide. It is shown that oleate can bind to metmyoglobin and azidometmyoglobin, but not oxymyoglobin. Catalase reduces the oxidation rates of oxymyoglobin in the presence or in the absence of oleate, substantiating the involvement of H2O2. The results are discussed in relation to the potential increase in tissue peroxidations in the presence of ischaemically elevated fatty acid concentrations.  相似文献   

9.
NAD(P)H acts as a two-electron reductant in physiological, enzyme-controlled processes. Under nonenzymatic conditions, a couple of one-electron oxidants easily oxidize NADH to the NAD(.) radical. This radical reduces molecular oxygen to the superoxide radical (O-(2)) at a near to the diffusion-controlled rate, thereby subsequently forming hydrogen peroxide (H(2)O(2)). Because peroxynitrite can act as a one-electron oxidant, the reaction of NAD(P)H with both authentic peroxynitrite and the nitric oxide ((. )NO) and O-(2) releasing compound 3-morpholinosydnonimine N-ethylcarbamide (SIN-1) was studied. Authentic peroxynitrite oxidized NADH with an efficiency of approximately 25 and 8% in the absence and presence of bicarbonate/carbon dioxide (HCO(3)(-)/CO(2)), respectively. NADH reacted 5-100 times faster with peroxynitrite than do the known peroxynitrite scavengers glutathione, cysteine, and tryptophan. Furthermore, NADH was found to be highly effective in suppressing peroxynitrite-mediated nitration reactions even in the presence of HCO(3)(-)/CO(2). Reaction of NADH with authentic peroxynitrite resulted in the formation of NAD(+) and O-(2) and, thus, of H(2)O(2) with yields of about 3 and 10% relative to the added amounts of peroxynitrite and NADH, respectively. Peroxynitrite generated in situ from SIN-1 gave virtually the same results; however, two remarkable exceptions were recognized. First, the efficiency of NADH oxidation increased to 60-90% regardless of the presence of HCO(3)(-)/CO(2), along with an increase of H(2)O(2) formation to about 23 and 35% relative to the amounts of added SIN-1 and NADH. Second, and more interesting, the peroxynitrite scavenger glutathione (GSH) was needed in a 75-fold surplus to inhibit the SIN-1-dependent oxidation of NADH half-maximal in the presence of HCO(3)(-)/CO(2). Similar results were obtained with NADPH. Hence, peroxynitrite or radicals derived from it (such as, e.g. the bicarbonate radical or nitrogen dioxide) indeed oxidize NADH, leading to the formation of NAD(+) and, via O-(2), of H(2)O(2). When peroxynitrite is generated in situ in the presence of HCO(3)(-)/CO(2), i.e. under conditions mimicking the in vivo situation, NAD(P)H effectively competes with other known scavengers of peroxynitrite.  相似文献   

10.
Evidence that hydroxyl radicals mediate auxin-induced extension growth   总被引:17,自引:0,他引:17  
Schopfer P  Liszkay A  Bechtold M  Frahry G  Wagner A 《Planta》2002,214(6):821-828
Reactive oxygen intermediates, i.e. the superoxide radical (O*-)(2), hydrogen peroxide (H2O2) and the hydroxyl radical (*OH), are generally regarded as harmful products of oxygenic metabolism causing cell damage in plants, animals and microorganisms. However, oxygen radical chemistry may also play a useful role in polymer breakdown leading to wall loosening during extension growth of plant cells controlled by the phytohormone auxin. Backbone cleavage of cell wall polysaccharides can be accomplished in vitro by (*OH) produced from H2O2 in a Fenton reaction or in a reaction catalyzed by peroxidase supplied with O2 and NADH. Here, we show that coleoptile growth of maize seedlings is accompanied by the release of reactive oxygen intermediates in the cell wall. Auxin promotes release of (O*-)(2) and subsequent generation of (*OH)when inducing elongation growth. Experimental generation of (*OH) in the wall causes an increase in wall extensibility in vitro and replaces auxin in inducing growth. Auxin-induced growth can be inhibited by scavengers of (O*-)(2), H2O2 or (*OH), or inhibitors interfering with the formation of these molecules in the cell wall. These results provide the experimental background for a novel hypothesis on the mechanism of plant cell growth in which (*OH), produced from (O*-)(2) and H2O2 by cell wall peroxidase, acts as a wall-loosening agent.  相似文献   

11.
Incubation of horse-heart oxymyoglobin or metmyoglobin with excess H2O2 causes formation of myoglobin(IV), followed by haem degradation. At the time when haem degradation is observed, hydroxyl radicals (.OH) can be detected in the reaction mixture by their ability to degrade the sugar deoxyribose. Detection of hydroxyl radicals can be decreased by transferrin or by OH scavengers (mannitol, arginine, phenylalanine) but not by urea. Neither transferrin nor any of these scavengers inhibit the haem degradation. It is concluded that intact oxymyoglobin or metmyoglobin molecules do not react with H2O2 to form OH detectable by deoxyribose, but that H2O2 eventually leads to release of iron ions from the proteins. These released iron ions can react to form OH outside the protein or close to its surface. Salicylate and the iron chelator desferrioxamine stabilize myoglobin and prevent haem degradation. The biological importance of OH generated using iron ions released from myoglobin by H2O2 is discussed in relation to myocardial reoxygenation injury.  相似文献   

12.
The new compound trimethylene-N(6),N(6')-bisadenine (L), in which two adenine molecules are linked together by a trimethylene bridge that connects the N(6) atoms, has been prepared. Reaction of L with HgCl(2) and ZnCl(2) in concentrated HCl solution leads to crystalline solids. The X-ray characterisation of the Hg(II) complex (H(2)L)[HgCl(4)].3H(2)O reveals that it is an outer-sphere complex in which the ligand is protonated at N(1) and N(1'). In contrast, the structure of the complex [H(2)L(ZnCl(3))(2)].2H(2)O shows the ligand co-ordinated to two different Zn(II) ions through the N(7) of both adenine fragments, the protons being located on the N(1) atoms. The latter compound constitutes the first crystallographic evidence of an inner sphere complex with bis-adenines and, for this reason, an equilibrium study was carried out on the Zn(II)-L-H(+) system. Potentiometric studies indicate that L is protonated in aqueous solution to form HL(+) and H(2)L(2+) with logK(H) values of 4.42 and 3.35 (25 degrees C, 0.10 M KNO(3)). The data from potentiometric titrations in the presence of Zn(2+) can be analysed considering the formation of the species LZn(2+), HLZn(3+), LZn(2)(4+) and HLZn(2)(5+), whose stability constants exceed the value expected for a monodentate interaction of the metal ion with adenine and suggest the possibility of a polydentate behaviour of L in the pH range 2.5-5.0. In contrast, spectrophotometric titrations carried out under conditions similar to those used in the synthetic work (1 M HCl) can be fitted with a model involving exclusively the H(2)LZn(4+) and H(2)LZn(2)(6+) species with logK(M) values reasonable for the interaction of Zn(II) with the N(7) of the protonated adenine fragments. Despite the H(2)LZn(2)(6+) species has a low stability, the spectrophotometric results are in agreement with its formation under the conditions in which the solid complex was prepared.  相似文献   

13.
The reaction of trans-[Ru(NH(3))(4)P(OEt)(3)NO](3+) and mitochondria was investigated through differential pulse polarography and fluorimetry. The nitrosyl complex undergoes one-electron reduction centered on the NO ligand site. The reaction between the mitochondrial reductor and trans-[Ru(NH(3))(4)P(OEt)(3)NO](3+) exhibits a second order specific rate constant calculated as k=2 x 10(1) M(-1) s(-1). The reduced species, trans-[Ru(NH(3))(4)P(OEt)(3)NO](2+), quickly releases NO, yielding trans-[Ru(NH(3))(4)P(OEt)(3)H(2)O](2+). The low toxicities of both trans-[Ru(NH(3))(4)P(OEt)(3)(NO)](2+) and trans-[Ru(NH(3))(4)P(OEt)(3)H(2)O](2+) and its ability to release NO after reductive activation in a biological medium make the nitrosyl compound a useful model of a hypotensive drug.  相似文献   

14.
Nitric oxide as an antioxidant.   总被引:21,自引:0,他引:21  
Benzoate monohydroxy compounds, and in particular salicylate, were produced during interaction of ferrous complexes with hydrogen peroxide (Fenton reaction) in a N2 environment. These reactions were inhibited when Fe complexes were flushed, prior to the addition in the model system, by nitric oxide. Methionine oxidation to ethylene by Fenton reagents was also inhibited by nitric oxide. Myoglobin in several forms such as metmyoglobin, oxymyoglobin, and nitric oxide-myoglobin were interacted with an equimolar concentration of hydrogen peroxide. Spectra changes in the visible region and the changes in membrane (microsomes) lipid peroxidation by the accumulation of thiobarbituric acid-reactive substances (TBA-RS) were determined. The results showed that metmyoglobin and oxymyoglobin were activated by H2O2 to ferryl myoglobin, which initiates membrane lipid peroxidation; but not nitric oxide-myoglobin, which, during interaction with H2O2, did not form ferryl but metmyoglobin which only poorly affected lipid peroxidation. It is assumed that nitric oxide, liganded to ferrous complexes, acts to prevent the prooxidative reaction of these complexes with H2O2.  相似文献   

15.
Extracellularly secreted plant peroxidases (POXs) are considered to catalyze the generation of reactive oxygen species (ROS) coupled to oxidation of plant hormone indole-3-acetic acid (IAA) and defense-related compounds salicylic acid (SA), aromatic monoamines (AMAs) and chitooligosaccharides (COSs). This review article consists of two parts, which describe H(2)O(2)-dependent and H(2)O(2)-independent mechanisms for ROS generation, respectively. Recent studies have shown that plant POXs oxidize SA, AMAs and COSs in the presence of H(2)O(2) via a conventional POX cycle, yielding the corresponding radical species, such as SA free radicals. These radical species may react with oxygen, and superoxide (O(2)(.-)) is produced. Through the series of reactions 2 moles of O(2)(.-) can be formed from 1 moles of H(2)O(2), thus leading to oxidative burst. It has been revealed that the ROS induced by SA, AMAs and COSs triggers the increase in cytosolic Ca(2+) concentration. Actually POXs transduce the extracellular signals into the redox signals that eventually stimulate the intracellular Ca(2+) signaling required for induction of defense responses. On the other hand, IAA can react with oxygen and plant POXs in the absence of H(2)O(2), by forming the ternary complex enzyme-IAA-O(2), which readily dissociates into enzyme, IAA radicals and O(2)(.-). This article covers the recent reports showing that extracellularly produced hydroxy radicals derived from O(2)(.-) mediate the IAA-induced cell elongation. Here a novel model for IAA signaling pathway mediated by extracellular ROS produced by cell-wall POXs is proposed. In addition, possible controls of the IAA-POX reactions by a fungal alkaloid are discussed.  相似文献   

16.
t-Butyl hydroperoxide and H2O2-Fe(2+)-EDTA-glutathione system which produces hydroxyl radicals did not affect the 15-hydroxy prostaglandin dehydrogenase activity in rabbit kidney cortex. On the other hand, H2O2-Fe(2+)-EDTA-glutathione system inhibited the prostaglandin delta 13 reductase activity. Mannitol, a scavenger of hydroxyl radicals, had no effect on the inhibitory action of this system, indicating that the effect of H2O2-Fe(2+)-EDTA-glutathione system on the prostaglandin delta 13 reductase may not be due to produced hydroxyl radicals. As a result of further investigation, it was shown that glutathione disulfide, which is synthesized concomitantly with hydroxyl radicals from H2O2-Fe(2+)-EDTA-glutathione, inhibited the prostaglandin delta 13 reductase activity. These results suggest that hydroperoxides and hydroxyl radicals may not be likely candidates for the modulator of the catabolism of prostaglandins in the kidney cortex, and that glutathione disulfide has the potential to modulate the prostaglandin catabolism by affecting the prostaglandin delta 13 reductase activity.  相似文献   

17.
Vanadium K-edge X-ray absorption spectroscopy (XAS) has been used to track the uptake and fate of VO(2+) ion in blood cells from Ascidia ceratodes, following exposure to dithiothreitol (DTT) or to DTT plus VO(2+). The full range of endogenous vanadium was queried by fitting the XAS of blood cells with the XAS spectra of model vanadium complexes. In cells exposed only to DTT, approximately 0.4% of a new V(III) species was found in a site similar to Na[V(edta)(H(2)O)]. With exposure to DTT and VO(2+), average intracellular [VO(aq)](2+) increased from 3% to 5%, and 6% of a new complexed form of vanadyl ion appeared evidencing a ligand array similar to [VO(edta)](2-). At the same time, the relative ratio of blood cell [V(H(2)O)(6)](3+) increased at the expense of [V(H(2)O)(5)(SO(4))](+) in a manner consistent with a significant increase in endogenous acidity. In new UV/Visible experiments, VO(2+) could be reduced to 7-coordinate [V(nta)(H(2)O)(3)] or [V(nta)(ida)](2-) with cysteine methyl ester in pH 6.5 solution. Ascorbate reduced [VO(edta)](2-) to 7-coordinate [V(edta)(H(2)O)](-), while [VO(trdta)](2-) was unreactive. These results corroborate the finding that the reductive EMF of VO(2+) is increased by the availability of a 7-coordinate V(III) product. Finally, a new and complete hypothesis is proposed for an ascidian vanadate reductase. The structure of the enzyme active site, the vanadate-vanadyl-vanadic reduction mechanism, the cellular locale, and elements of the regulatory machinery governing the biological reduction of vanadate and vanadyl ion by ascidians are all predicted. Together these constitute the new field of vanadium redox enzymology.  相似文献   

18.
The DNA-binding proteins from starved cells (Dps) are a family of proteins induced in microorganisms by oxidative or nutritional stress. Escherichia coli Dps, a structural analog of the 12-subunit Listeria innocua ferritin, binds and protects DNA against oxidative damage mediated by H(2)O(2). Dps is shown to be a Fe-binding and storage protein where Fe(II) oxidation is most effectively accomplished by H(2)O(2) rather than by O(2) as in ferritins. Two Fe(2+) ions bind at each of the 12 putative dinuclear ferroxidase sites (P(Z)) in the protein according to the equation, 2Fe(2+) + P(Z) --> [(Fe(II)(2)-P](FS)(Z+2) + 2H(+). The ferroxidase site (FS) bound iron is then oxidized according to the equation, [(Fe(II)(2)-P](FS)(Z+2) + H(2)O(2) + H(2)O --> [Fe(III)(2)O(2)(OH)-P](FS)(Z-1) + 3H(+), where two Fe(II) are oxidized per H(2)O(2) reduced, thus avoiding hydroxyl radical production through Fenton chemistry. Dps acquires a ferric core of approximately 500 Fe(III) according to the mineralization equation, 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2Fe(III)OOH((core)) + 4H(+), again with a 2 Fe(II)/H(2)O(2) stoichiometry. The protein forms a similar ferric core with O(2) as the oxidant, albeit at a slower rate. In the absence of H(2)O(2) and O(2), Dps forms a ferrous core of approximately 400 Fe(II) by the reaction Fe(2+) + H(2)O + Cl(-) --> Fe(II)OHCl((core)) + H(+). The ferrous core also undergoes oxidation with a stoichiometry of 2 Fe(II)/H(2)O(2). Spin trapping experiments demonstrate that Dps greatly attenuates hydroxyl radical production during Fe(II) oxidation by H(2)O(2). These results and in vitro DNA damage assays indicate that the protective effect of Dps on DNA most likely is exerted through a dual action, the physical association with DNA and the ability to nullify the toxic combination of Fe(II) and H(2)O(2). In the latter process a hydrous ferric oxide mineral core is produced within the protein, thus avoiding oxidative damage mediated by Fenton chemistry.  相似文献   

19.
Aldehyde dehydrogenase 2 (ALDH2) catalyzes oxidation of toxic aldehydes to carboxylic acids. Physiologic levels of Mg(2+) ions influence ALDH2 activity in part by increasing NADH binding affinity. Traditional fluorescence measurements monitor the blue shift of the NADH fluorescence spectrum to study ALDH2-NADH interactions. By using time-resolved fluorescence spectroscopy, we have resolved the fluorescent lifetimes (τ) of free NADH (τ=0.4 ns) and bound NADH (τ=6.0 ns). We used this technique to investigate the effects of Mg(2+) on the ALDH2-NADH binding characteristics and enzyme catalysis. From the resolved free and bound NADH fluorescence signatures, the K(D) for NADH with ALDH2 ranged from 468 μM to 12 μM for Mg(2+) ion concentrations of 20 to 6000 μM, respectively. The rate constant for dissociation of the enzyme-NADH complex ranged from 0.4s(-1) (6000 μM Mg(2+)) to 8.3s(-1) (0 μM Mg(2+)) as determined by addition of excess NAD(+) to prevent re-association of NADH and resolving the real-time NADH fluorescence signal. The apparent NADH association/re-association rate constants were approximately 0.04 μM(-1)s(-1) over the entire Mg(2+) ion concentration range and demonstrate that Mg(2+) ions slow the release of NADH from the enzyme rather than promoting its re-association. We applied NADH fluorescence lifetime analysis to the study of NADH binding during enzyme catalysis. Our fluorescence lifetime analysis confirmed complex behavior of the enzyme activity as a function of Mg(2+) concentration. Importantly, we observed no pre-steady state burst of NADH formation. Furthermore, we observed distinct fluorescence signatures from multiple ALDH2-NADH complexes corresponding to free NADH, enzyme-bound NADH, and, potentially, an abortive NADH-enzyme-propanal complex (τ=11.2 ns).  相似文献   

20.
《Free radical research》2013,47(6):415-422
Incubation of horse-heart oxymyoglobin or metmyoglobin with excess H2O2 causes formation of myoglobin(IV), followed by haem degradation. At the time when haem degradation is observed, hydroxyl radicals (.OH) can be detected in the reaction mixture by their ability to degrade the sugar deoxyribose. Detection of hydroxyl radicals can be decreased by transferrin or by OH scavengers (mannitol, arginine, phenylalanine) but not by urea. Neither transferrin nor any of these scavengers inhibit the haem degradation. It is concluded that intact oxymyoglobin or metmyoglobin molecules do not react with H2O2 to form OH detectable by deoxyribose, but that H2O2 eventually leads to release of iron ions from the proteins. These released iron ions can react to form OH outside the protein or close to its surface. Salicylate and the iron chelator desferrioxamine stabilize myoglobin and prevent haem degradation. The biological importance of OH generated using iron ions released from myoglobin by H2O2 is discussed in relation to myocardial reoxygenation injury.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号