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.
Crocin in aqueous solution is oxidized by ferrylmyoglobin, MbFe(IV)=O, in a second order reaction with k = 183 1 · mol-1 · s-1, AH298 = 55.0 kJ · mol-1, and ΔLS298 = -17 J · mol-1 K-1 (pH = 6.8, ionic strength 0.16 (NaCl), 25°C), as studied by stopped-flow spectroscopy. The reaction has 1:1 stoichiometry to yield metmyoglobin, MbFe(III), and has AGo = -11 kJ · mol-1, as calculated from the literature value E0 = +0.85 V (pH = 7.4) vs. NHE for MbFe(IV)=O/MbFe(III) and from the half-peak potential +0.74 V (vs. NHE in aqueous 0.16 NaCl, pH = 7.4) determined by cyclic voltammetry for the one-electron oxidation product of crocin, for which a cation radical structure is proposed and which has a half-peak potential of +0.89 V for its formation from the two-electron oxidation product of crocin. The fer-rylmyoglobin protein-radical, MbFe(IV)=O, reacts with crocin with 2:l stoichiometq to yield MbFe(IV)= 0, as determined by ESR spectroscopy, in a reaction faster than the second order protein-radical generating reaction between H2O2 and MbFe(III), for which latter reaction k = 137 L · mol-1 · s-1, ΔH298 = 51.5 kJ · mol-1, and ΔH298 = -31 J · mol-1 · K-1 (pH = 6.8, ionic strength = 0.16 (NaCI), 25°C) was determined. Based on the difference between the stoichiometry for the reaction between crocin and each of the two hypervalent forms of myoglobin, it is concluded in agreement with the determined half peak reduction potentials, that the crocin cation radical is less reducing compared to crocin, as the cation radical can reduce the protein radical but not the iron(IV) centre in hypervalent myoglobin.  相似文献   

2.
Tetrazolium Blue (TBCl2) and Nitrotetrazolium Blue (NTBCl2) cathodically reduced in non aqueous solvents form radicals with the center of unpaired electron on the tetrazolyl ring (TBH2, NTBH2) as detected by EPR spectroscopy. After prolonged reduction, formazans (TBH2, NTBH2) are formed and are then further reduced to the nitro-centered anion radical (from NTBH2) and the azogroup-centered anion radical (from TBH2). The first cathodic peak in the cyclovoltammetric study in the region from -0.3 to -0.6 V vs. SCE (saturated calomel electrode) is irreversible and indicates an adsorption and diffusion process on the platinum and mercury electrodes. Formation of TBH+ and NTBH+ is assumed. The second peak, in the region from -0.8 to -1.3 V vs. SCE, is nearly reversible and coupled with the formation of TBH- and NTBH- radicals. UV-VIS spectra measured during the reduction show isosbestic points at the conversions: TB++ → TBH+, NTB++ - NTBH+; further, TBH+ → TBH2 and NTBH+ → NTBH2. The characteristic colours of the solutions observed can be used to characterise the reduction state of tetrazolium salts.  相似文献   

3.
Abstract: In this study we have examined (1) the integrated function of the mitochondrial respiratory chain by polarographic measurements and (2) the activities of the respiratory chain complexes I, II–III, and IV as well as the ATP synthase (complex V) in free mitochondria and synaptosomes isolated from gerbil brain, after a 30-min period of graded cerebral ischaemia. These data have been correlated with cerebral blood flow (CBF) values as measured by the hydrogen clearance technique. Integrated functioning of the mitochondrial respiratory chain, using both NAD-linked and FAD-linked substrates, was initially affected at CBF values of ∼35 ml 100 g−1 min−1, and declined further as the CBF was reduced. The individual mitochondrial respiratory chain complexes, however, showed differences in sensitivity to graded cerebral ischaemia. Complex I activities decreased sharply at blood flows below ∼30 ml 100 g−1 min−1 (mitochondria and synaptosomes) and complex II–III activities decreased at blood flows below 20 ml 100 g−1 min−1 (mitochondria) and 35–30 ml 100 g−1 min−1 (synaptosomes). Activities declined further as CBF was reduced below these levels. Complex V activity was significantly affected only when the blood flow was reduced below 15–10 ml 100 g−1 min−1 (mitochondria and synaptosomes). In contrast, complex IV activity was unaffected by graded cerebral ischaemia, even at very low CBF levels.  相似文献   

4.
The reaction of FeII oxalate with hydrogen peroxide and dioxygen was studed for oxalate concentrations up to 20 mM and pH 2-5, under which conditions mono- and bis-oxalate comlexes (FeII(ox) and FeII(ox)22-) and uncomplexed Fe2+ must be considered. The reaction of FeII oxalate with hydrogen peroxide (Fe2+ + H2O2 → Fe3+ + *OH + OH-) was monitored in continuous flow by ESR with t-butanol as a radical trap. The reaction is much faster than for uncomplexed Fe2+ and a rate constant, k = 1 × 104 M-1 s-1 is deduced for FeII(ox). The reaction of FeII oxalate with dioxygen is strongly pH dependent in a manner which indicates that the reactive species is FeII(ox)22-, for which an apparent second order rate constant, k = 3.6 M-1 s-1, is deduced. Taken together, these results provide a mechanism for hydroxyl radical production in aqueous systems containing FeII complexed by oxalate. Further ESR studies with DMPO as spin trap reveal that reaction of FeII oxalate with hydrogen peroxide can also lead to formation of the carboxylate radical anion (CO2*-), an assignment confirmed by photolysis of FeIII oxalate in the presence of DMPO.  相似文献   

5.
Nitric oxide (NO) is a highly reactive, membrane-permeable free radical, which has recently emerged as an important antioxidant. Here we investigated the protective effect of NO against the toxicity and NH4+ accumulation in rice leaves caused by excess CuSO4 (10 mmol L−1). It was found that free radical scavengers (sodium benzoate, thiourea, and reduced glutathione) reduced the toxicity and NH4+ accumulation in rice leaves caused by excess CuSO4. NO donor sodium nitroprusside (SNP) was also effective in reducing CuSO4-induced toxicity and NH4+ accumulation in rice leaves. The protective effect of SNP on the toxicity and NH4+ accumulation can be reversed by 2-(4-carboxy-2-phenyl)-4,4,5,5-tetramethyl- imidazoline-1-oxyl-3-oxide, a NO scavenger, suggesting that the protective effect of SNP is attributable to NO released. Results obtained in the present study suggest that reduction of CuSO4-induced toxicity and NH4+ accumulation by SNP is most likely mediated through its ability to scavenge active oxygen species.  相似文献   

6.
Adriamycin free radical anion concentrations have been correlated with production of 7-deoxyadriamycin aglycone in a reaction catalyzed by NADPH-cytochrome c reductase. The free radical species is detected by electron spin resonance (ESR) spectroscopy and quantified by double integrations. The 7-deoxyaglycone product is isolated by thin-layer chromatography (TLC) and quantified by fluorometry. As the concentration of adriamycin increases, a concomitant increase in aglycone and free radical levels occurs. These results as well as those with inhibitors Vitamin K3, Vitamin E, and 5,5-dimethyl-1-pyrroline-1-oxide (DMPO) point to an obligatory free radical intermediate in the metabolism of adriamycin. DMPO inhibits the reaction under aerobic conditions only, and shows no effect under anaerobiosis at the concentrations studied here. Vitamin E and aerobic DMPO act as free radical scavangers, while Vitamin K3 competes for the reducing power of NADPH in the NADPH-cytochrome c reductase system.  相似文献   

7.
Acetylene was reduced by zinc amalgam in the presence of three synthetic polynuclear complexes: {[Mg2Mo8O22(OMe)6(MeOH)4]−2·[Mg(MeOH)6]2+}6MeOH (I), (Bu4N)2[Fe4S4(SPh)4] (II), [Me4N][VFe3S4Cl3(DMF)3]·2DMF (III) and the iron-molybdenum cofactor of nitrogenase Azotobacter vinelandii MoFe7(S2−)9·homocitrate, FeMo-co (IV). Thiophenol was found to greatly facilitate the reaction in the presence of complexes I, II, IV. The reaction is catalytic and for I and IV proceeds at the amalgam surface. Thiophenol seems to increase the adsorption of the complexes, serving as an electron bridge to transfer electrons to the catalyst. In the case of II a homogeneous reduction of the substrate occurs presumably after the cluster reduction at the surface and with III the catalytic reduction proceeds only under the action of sodium amalgam; no thiophenol cocatalytic action is observed. Relevance to N2 enzymatic reduction is discussed.  相似文献   

8.
J. Butler  G.G. Jayson  A.J. Swallow 《BBA》1975,408(3):215-222

1. 1. The superoxide anion radical (O2) reacts with ferricytochrome c to form ferrocytochrome c. No intermediate complexes are observable. No reaction could be detected between O2 and ferrocytochrome c.

2. 2. At 20 °C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4 · 106 M−1 · s−1 and as the pH increases above 6.7 the rate constant steadily decreases. The dependence on pH is the same for tuna heart and horse heart cytochrome c. No reaction could be demonstrated between O2 and the form of cytochrome c which exists above pH ≈ 9.2. The dependence of the rate constant on pH can be explained if cytochrome c has pKs of 7.45 and 9.2, and O2 reacts with the form present below pH 7.45 with k = 1.4 · 106 M−1 · s−1, the form above pH 7.45 with k = 3.0 · 105 M−1 · s−1, and the form present above pH 9.2 with k = 0.

3. 3. The reaction has an activation energy of 20 kJ mol−1 and an enthalpy of activation at 25 °C of 18 kJ mol−1 both above and below pH 7.45. It is suggested that O2 may reduce cytochrome c through a track composed of aromatic amino acids, and that little protein rearrangement is required for the formation of the activated complex.

4. 4. No reduction of ferricytochrome c by HO2 radicals could be demonstrated at pH 1.2–6.2 but at pH 5.3, HO2 radicals oxidize ferrocytochrome c with a rate constant of about 5 · 105–5 · 106 M−1 · s−1

.  相似文献   


9.
Five heterometallic compounds with formulae [Ba(H2O)4Cr2(μ-OH)2(nta)2] · 3H2O (I), [M(bpy)2(H2O)2] [Cr2(OH)2(nta)2] · 7H2O, where M2+ = Zn, (II); Ni, (III); Co, (IV) and [Mn(H2O)3(bpy)Cr2(OH)2(nta)2] · (bpy) · 5H2O (V); bpy = 2,2′-bipyridine, (nta = nitrilotriacetate ion) have been prepared by reaction of I with the corresponding MII-sulfates in the presence of 2,2′-bipyridine. Substances I–V have been characterized by magnetic susceptibility measurements, EPR and X-ray determinations. I represents a 2D coordination polymer formed by coordination of centrosymmetrical dimeric chromium(III) units and Barium cations. The 10-coordinate Ba polyhedron is completed by four water molecules. Compounds II–IV are isostructural and consist of non-centrosymmetric dimeric anions [Cr2(μ-OH)2(nta)2]2−, complex cations [MII(bpy)2(H2O)2]2+ and solvate water molecules. The octahedral coordination of chromium atoms implies four donor atoms of the nta3− ligands and two bridging OH groups. Multiple hydrogen bonds of coordinated and solvate water molecules link anions and cations in a 3D network. A similar [Cr2(μ-OH)2(nta)2]2− unit is found in V. The bridging function is performed by a carboxylate oxygen atom of the nta ligand that leads to the formation of a trinuclear complex [Mn(bpy)(H2O)2Cr2(μ-OH)2(nta)2]. Experimental and calculated frequency and temperature dependences of EPR spectra of these compounds are presented. The fine structure appearing on the EPR spectra of compound V is analyzed in detail at different temperatures. It is established that the main part of the EPR signals is due to the transitions in the spin states of a spin multiplet with S = 2. Analyses of experimental and calculated spectra confirm the absence of interaction between metal ions (MII) and Cr-dimers in complexes III and IV and the presence of weak Mn–Cr interactions in V. The temperature dependence of magnetic susceptibilities for I–V was fitted on the basis of the expression derived from isotropic Hamiltonian including a bi-quadratic exchange term.  相似文献   

10.
σ-Methyl-(η5-indenyl) chromium tricarbonyl (III) rearranges quantitatively into η6-1-endo-methylindene) chromium tricarbonyl (IV) in C6D6 solution at 30–60°C. Methyl group attachment to the positions 2 or 3 of indenyl ligand in (III) has no influence on the activation parameters of this ricochet inter-ring haptotropic rearrangement (ΔG#=23.6 kcal mol−1; ΔH#=18.9±0.2 kcal mol−1; ΔS#=−18.6±0.2 cal K−1 mol−1). (IV) undergoes further irreversible isomerization at 60–120° into (ν6-3-methylindene) chromium tricarbonyl (V) with a higher activation barrier (ΔG#=28.5±0.1 kcal mol−1) via two consecutive [1,5]-sigmatropic hydrogen shifts. The mechanisms of both rearrangements have been studied in detail using density functional theory (DFT) calculations with extended basis sets. Calculations show that the rearrangement (III) → (IV) proceeds in two steps. Methyl group migration from chromium into position 1 of the indenyl ligand is the rate-determining step leading to the formation of the 16-electron intermediate (VII). The calculated activation barrier (Ea=19.6 kcal mol−1) is in good agreement with the experimental one. Further rearrangement (VII) → (V) proceeds via a trimethylenemethane-type transition state (XVIII) with an activation barrier 11.8 kcal mol−1. The coordination of the chromium tricarbonyl group at the six-membered ring has only minor influence on the kinetic parameters of the hydrogen [1,5]-sigmatropic shift in indene.  相似文献   

11.
The interaction of superoxide ion and ascorbate anion with anthracycline antibiotics (adriamycin and aclacinimycin A) as well as with their Fe3+ complexes has been studied in aprotic and protic media. It was found that both superoxide and ascorbate reduce anthracyclines to deoxyaglycons via a one-electron transfer mechanism under all conditions studied. The reaction of ascorbate anion with adriamycin and aclacinomycin A in aqueous solution proceeded only in the presence of Fe3+ ions; it is supposed that an active catalytic species was Fe3+ adriamycin. It is also supposed that the reduction of anthracycline antibiotics by O,7 and ascorbate in cells may increase their anticancer effect.  相似文献   

12.
Nitric oxide, a gaseous free radical, is poorly reactive with most biomolecules but highly reactive with other free radicals. Its ability to scavenge peroxyl and other damaging radicals may make it an important antioxidant in vivo, particular in the cardiovascular system, although this ability has been somewhat eclipsed in the literature by a focus on the toxicity of peroxynitrite, generated by reaction of O·-2 with NO· (or of NO- with O2). On balance, experimental and theoretical data support the view that ONOO- can lead to hydroxyl radical (OH·) generation at pH 7.4, but it seems unlikely that OH· contributes much to the cytotoxicity of ONOO-. The cytotoxicity of ONOO- may have been over-emphasized: its formation and rapid reaction with antioxidants may provide a mechanism of using NO· to dispose of excess O·-2, or even of using O·-2 to dispose of excess NO·, in order to maintain the correct balance between these radicals in vivo. Injection or instillation of “bolus” ONOO- into animals has produced tissue injury, however, although more experiments generating ONOO- at steady rates in vivo are required. The presence of 3-nitrotyrosine in tissues is still frequently taken as evidence of ONOO- generation in vivo, but abundant evidence now exists to support the view that it is a biomarker of several “reactive nitrogen species”. Another under-addressed problem is the reliability of assays used to detect and measure 3-nitrotyrosine in tissues and body fluids: immunostaining results vary between laboratories and simple HPLC methods are susceptible to artefacts. Exposure of biological material to low pH (e.g. during acidic hydrolysis to liberate nitrotyrosine from proteins) or to H2O2 might cause artefactual generation of nitrotyrosine from NO-2 in the samples. This may be the origin of some of the very large values for tissue nitrotyrosine levels quoted in the literature. Nitrous acid causes not only tyrosine nitration but also DNA base deamination at low pH: these events are relevant to the human stomach since saliva and many foods are rich in nitrite. Several plant phenolics inhibit nitration and deamination in vitro, an effect that could conceivably contribute to their protective effects against gastric cancer development.  相似文献   

13.
Fourteen flavonoid aglycones, and the flavonoid glyco-side rutin, with redox potentials ranging from 0.20 (myricetin) to 0.83 V (chrysin) vs. NHE, as determined by cyclic voltammetry at 23°C in aqueous 50 mM phosphate, ionic strength 0.16 (NaCI) with pH = 7.4 and compared with redox potentials determined for four cinnamic acid derivatives, were all found to reduce ferrylmyoglobin, MbFe(IV)=O, to metmyoglo-bin, MbFe(III). Reaction stoichiometry depends strongly on the number of hydroxyl groups in the flavonoid B-ring. All compounds with 3',4'-dihydroxy substitution reduce 2 equivalents of MbFe(IV)=O, whereas naringenin, hesperitin and kaempferol, with one hydroxyl group in the B-ring, reduce with a one-to-one stoichiometry. As studied spectrophotometrically under pseudo-first-order conditions with flavonoids in excess, rutin and apigenin react with MbFe(IV)=O with very similar and moderately high activation enthalpies of ΔH‡298 = 69 ± 1kJ mol-1 and ΔH‡298 = 65 ± 3kJ mol-1, respectively, and with positive activation entropies of ΔH‡298 = 23 ± 4Jmol-1 K-1 and ΔS‡298 = 13 ± 9Jmol-1K-1, respectively, in agreement with outer-sphere electron transfer as rate determining. For the fifteen plant polyphenols only qualitative relations exist between redox potential and rate constants rather than a linear free energy relationship (r2 = 0.503), and especially the flavone apigenin was found more efficient as reducing agent. For the flava-nones, a linear relation (r2 = 0.971) indicate that, in the absence of a 2,3 double bond, removal of the 4-carbonyl group or addition of a 3-hydroxy group only has minor effect on reactivity. The flavonols are the most efficient reducing agents, effectively reducing MbFe(IV)=O to MbFe(III) and establishing a steady state distribution between the flavonol and MbFe(III) and oxymyoglobin, MbFe(II)O2. Oxidised flavonols reduces MbFe(III) to MbFe(II)02 very efficiently and much faster than the parent flavonol.  相似文献   

14.
The reaction of the superoxide radical anion (O2), with the semi-oxidized tryptophan neutral radical (Trp·) generated from tryptophan (Trp) by pulse radiolysis has been observed in a variety of functionalized Trp derivatives including peptides. It is found that the reaction proceeds 4-5 times faster in positively charged peptides, such as Lys-Trp-Lys, Lys-Gly-Trp-Lys and Lys-Gly-Trp-Lys-O-tert-butyl, than in solutions of the negatively charged N-acetyl tryptophan (NAT). However, the reactivity of O2 with the Trp· radical is totally inhibited upon binding of these peptides to micelles of negatively charged SDS and is reduced upon binding to native DNA. By contrast, no change in reactivity is observed in a medium containing CTAB, where the peptides cannot bind to the positively charged micelles. On the other hand, the reactivity of the Trp· radical formed from NAT with O2 is reduced to half that of the free Trp· in buffer but is markedly increased in CTAB micelles. The models studied here incorporate elements of the complex environment in which Trp· and O2 may be concomitantly formed in biological system and demonstrate the magnitude of the influence such elements may have on the kinetics of reactions involving these two species.  相似文献   

15.
The nicotinamide adenine dinucleotide dimers (NAD)2 obtained by electrochemical reduction of NAD+ are oxidized by adriamycin in anaerobic photocatalyzed reaction yielding NAD+ and 7-deoxyadriamyci-none. Under the same conditions NADH is not oxidized.  相似文献   

16.
目的:探讨阿霉素对口腔鳞癌干细胞迁移、侵袭、凋亡的影响及其可能的机制。方法:体外培养人口腔鳞癌细胞系SCC25,通过流式细胞术分选CD44-和CD44+细胞,RT-PCR检测CD44-和CD44+细胞的Oct4、CD133、CD44和GAPDH的m RNA表达;检测和比较CD44-和CD44+细胞的克隆形成能力。CD44+细胞用阿霉素或β-catenin抑制剂LF3进行处理,分别使用Transwell和细胞划痕检测细胞侵袭和迁移能力,一步法TUNEL检测细胞凋亡水平,WB检测β-catenin和TCF-4的蛋白表达。结果:流式细胞术成功分离CD44-和CD44+细胞,RT-PCR检测CD44+细胞高表达Oct4、CD133和CD44 m RNA,CD44-细胞弱表达Oct4m RNA,不表达CD133和CD44 m RNA;CD44+细胞的克隆形成能力显示显著强于CD44-细胞(P<0.05)。阿霉素显著降低了CD44+细胞的侵袭能力和迁移能力(P<0.05),显著提高了CD44+细胞的凋亡率(P<0.05);阿霉素显著降低了CD44+细胞β-catenin和TCF-4的蛋白表达(P<0.05),LF3对β-catenin和TCF-4蛋白表达的影响与阿霉素比较无显著差异(P>0.05)。结论:阿霉素可能通过抑制Wnt/β-catenin信号通路降低口腔鳞癌干细胞迁移、侵袭能力,促进细胞凋亡。  相似文献   

17.
Free radicals have been suggested to play a role in adriamycin-induced cardiomyopathy. Adriamycin-induced myocardial effects were examined in rats maintained on a vitamin E deficient diet. Animals were divided into four groups: I, control; II, adriamycin-treated; III, vitamin E deficient diet; IV, vitamin E deficient diet plus adriamycin treatment. Adriamycin-treated animals (groups II and IV) were given six injections (i.p.) over two weeks for producing a cumulative dose of 15 mg/kg. Animals in groups III and IV were placed on vitamin E deficient diet starting two weeks prior to the first injection of adriamycin or vehicle. Myocardial tissue analysis were performed on animals sacrificed 1 week after the last injection. Mortality was significantly higher in group IV which also showed doubling of myocardial malondialdehyde content relative to the non-adriamycin-treated vitamin E deficient group (III). Myocardial cell damage in group IV was characterized by separation of the external lamina, subsarcolemmal changes, mitochondrial swelling and myofibril dropout. Group II hearts showed only a mild dilation of the sarcotubules and swelling of the mitochondria. Total sialic acid content of the sarcolemma in groups II, III and IV was 55, 90 and 24% of the control values in group I. These data show a characteristic sarcolemmal injury produced by adriamycin in hearts of animals with reduced antioxidant capacity which is probably mediated by increased free radical activity as well as lipid peroxidation.  相似文献   

18.
1. Rate constants for reduction of paraquat ion (1,1′-dimethyl-4,4′-bipyridy-lium, PQ2+) to paraquat radical (PQ+·) by eaq and CO2· have been measured by pulse radiolysis. Reduction by eaq is diffusion controlled (k = 8.4·1010 M−1·s−1) and reduction by CO2· is also very fast k = 1.5·1010 M−1·s−1).

2. The reaction of paraquat radical with oxygen has been analysed to give rate constants of 7.7·108 M−1·s−1 and 6.5·108 M−1·s−1 for the reactions of paraquat radical with O2 and O2·, respectively. The similarity in these rate constants is in marked contrast to the difference in redox potentials of O2 and O2· (− 0.59 V and + 1.12 V, respectively).

3. These rate constants, together with that for the self-reaction of O2·, have been used to calculate the steady-state concentration of O2· under conditions thought to apply at the site of reduction of paraquat in the plant cell. On the basis of these calculations the decay of O2· appears to be governed almost entirely by its self-reaction, and the concentration 5 μm away from the thylakoid is still 90% of that at the thylakoid itself. Thus, O2· persists long enough to diffuse as far as the chloroplast envelope and tonoplast, which are the first structures to be damaged by paraquat treatment. O2· is therefore sufficiently long-lived to be a candidate for the phytotoxic product formed by paraquat in plants.  相似文献   


19.
Nitrogen dioxide (NO2) is a key biological oxidant. It can be derived from peroxynitrite via the interaction of nitric oxide with superoxide, from nitrite with peroxidases, or from autoxidation of nitric oxide. In this study, submicromolar concentrations of NO2 were generated in < 1 μs using pulse radiolysis, and the kinetics of scavenging NO2 by glutathione, cysteine, or uric acid were monitored by spectrophotometry. The formation of the urate radical was observed directly, while the production of the oxidizing radical obtained on reaction of NO2 with the thiols (the thiyl radical) was monitored via oxidation of 2,2′-azino-bis-(3-ethylthiazoline-6-sulfonic acid). At pH 7.4, rate constants for reaction of NO2 with glutathione, cysteine, and urate were estimated as 2 × 107, 5 × 107, and 2 × 107 M−1 s−1, respectively. The variation of these rate constants with pH indicated that thiolate reacted much faster than undissociated thiol. The dissociation of urate also accelerated reaction with NO2 at pH > 8. The thiyl radical from GSH reacted with urate with a rate constant of 3 × 107 M−1 s−1. The implications of these values are: (i) the lifetime of NO2 in cytosol is < 10 μs; (ii) thiols are the dominant ‘sink’ for NO2 in cells/tissue, whereas urate is also a major scavenger in plasma; (iii) the diffusion distance of NO2 is 0.2 μm in the cytoplasm and < 0.8 μm in plasma; (iv) urate protects GSH against depletion on oxidative challenge from NO2; and (v) reactions between NO2 and thiols/urate severely limit the likelihood of reaction of NO2 with NO• to form N2O3 in the cytoplasm.  相似文献   

20.
The absorption spectra of polyadenylic acid (polyA) radicals in N20 saturated aqueous solution have been measured as a function of time (up to 15 s) following an 0.4μS electron pulse. The spectra and their changes were analysed by comparison with those from monomeric adenine derivatives (nucleosides and nucleotides) which had been studied by Steenken.1

The reaction of OH· radicals with the adenine moiety in poly A results in the formation of two hvdroxvl adducts at the positions C-4 [polyA40H·] and C-8 [polyA80H·]. Each OH-adduct undergoes a unimol-ecular transformation reaction before any bimolecular or other unimolecular decay occurs. These reactions are characterized by different rate constants and pH dependencies. The polyA40H· adduct undergoes a dehydration reaction to yield a neutral N6 centered radical (rate constant Kdeh= 1.4 × 104s-1 at pH7.3). This reaction is strongly inhibited by H+. In comparison with the analogous reactions in adenosine phosphates, the kinetic pK value for its inhibition is two pH units higher. This shift is the result of the counter ion condensation or double-strand formation. The polyA80H· adduct undergoes an imidazole ring opening reaction to yield an enol type of formamidopyrimidine radical with the resulting base damage (kr.o. = 3.5 × 104 s -1 at pH7.3). This reaction in contrast is strongly catalysed by H+and OH-, similar as for adenosine but different compared to the nucleotides.  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