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 15 毫秒
1.
Protein radicals were selectively generated by reaction with azide radicals on Trp and Tyr residues in insulin, β-lactoglobulin, pepsin, chymotrypsin, and bovine serum albumin at rate constants in the range (2.9–19) × 108 M? 1 s? 1. Monohydrogen ascorbate reduced tryptophanyl radicals in chymotrypsin and pepsin with rate constants in the narrow range of (1.6–1.8) × 108 M? 1 s? 1, whereas β-lactoglobulin tryptophanyl radicals reacted almost 10 times slower. The corresponding values for the protein tyrosyl radicals were about an order of magnitude smaller. Comparison of the rate constants of reactions of free and protein-bound tryptophanyl and tyrosyl radicals showed that, in most cases, the location of the radicals in the protein chain did not constitute a major barrier to the reaction with monohydrogen ascorbate. The results suggest that, under physiological concentrations of dioxygen, monohydrogen ascorbate is likely to be a significant target of protein radicals. It seems likely, therefore, that reaction with protein radicals may be responsible for much of the well-documented loss of ascorbate in living organisms subjected to oxidative stress.  相似文献   

2.
Transient spectra and kinetic data of Tiron (1,2-dihydroxybenzene-3,5-disulphonic acid) are reported, obtained after pulse-radiolytic oxidation by hydroxyl radicals (°OH), superoxide anions (O2?) or a combination of both oxygen radicals. The rate constant with °OH radicals was determined at 1.0·109 M?1·s?1. Contrary to a previous report (Greenstock, C.L. and Miller, R.W. (1975) Biochim. Biophys. Acta 396, 11–16), the rate constant with O2? of 1.0·107 M?1·s?1 is lower by one order of magnitude; also the semiquinone absorbs at 300 nm rather than at 400 nm. The ratio of the rate constants with °OH and O2? of 100 again demonstrates that any oxidation reaction by the latter radical is unspecific due to the more efficient reaction of °OH radicals, leading to the same products with catechol compounds.  相似文献   

3.
《Free radical research》2013,47(4-6):289-294
Aroxyl radicals of fifteen structurally distinct flavonoids were generated by attack of azide radicals (N3) on the parent compounds dissolved in aqueous solution at pH 11.5. Generation rate constants were all found to be very high (2.4 – 8.8 × 109dm3mol?1s?1), whereas the decay rates differed considerably, ranging from 105 to 108dm3mol?1s?1. In most cases the spectral characteristics of the transient aroxyl radicals relate to structural features of the parent compounds and according to spectral similarities they can be classed in three distinct groups (with only two exceptions).

Although the data do not conclusively prove that the biological function of flavonoids might be the scavenging of radicals, the very high rate constants of formation and the relative stability of some of the aroxyl radicals, are in support of such a hypothesis.  相似文献   

4.
Neutral tryptophan (*Trp) and tyrosine (TyrO(*)) radicals are repaired by certain flavonoids in buffer, in micelles and in human serum albumin (HSA) with corresponding formation of semioxidized flavonoid radicals. In deaerated buffer, *Trp but not TyrO(*) radicals react with catechin. In micelles, quercetin and rutin repair both *Trp and TyrO(*) radicals. In addition to amino acid reactivity, microenvironmental factors and nature of the flavonoids govern this repair. Electron transfer efficiencies from quercetin to negatively charged *Trp radicals are 100% in the micellar pseudophases of positively charged cetyltrimethylammonium bromide, (CTAB), and neutral Triton X100 (TX100), but 55% in negatively charged sodium dodecyl sulfate (SDS). In oxygen-saturated CTAB micelles, quercetin also reacts with the superoxide radical anion. When bound to domain IIA of HSA, quercetin repairs, by intra- or intermolecular encounter, less than 20% of oxidative damage to HSA. Quercetin can also repair freely circulating oxidized molecules with repair efficiencies falling to 7% for oxidized *Trp, Tyr and alpha-MSH and to less than 2% for urate radical. This limited effectiveness is attributed both to the inaccessibility of bound quercetin and rutin toward radicals of circulating molecules and to the diffusion-controlled recombination of these radicals.  相似文献   

5.
Hiroshi Seki  Masashi Imamura 《BBA》1981,635(1):81-89
The reactions of ferrocytochrome c with Br?2, (SCN)?2, N3 and OH radicals were followed by measuring the change in the optical spectra of cytochrome c on γ-irradiation as well as the rate of change of absorbance upon pulse irradiation.Ferrocytochrome c is oxidized to ferricytochrome c by Br?2, (SCN)?2 or N3 radical with an efficiency of about 100% through a second-order process in which no intermediates were observed. The rate constants in neutral solutions at I = 0.073 are 9.7 · 108 M?1 · s?1, 7.9 · 108 M?1 · s?1, 1.3 · 109 M?1 · s?1 for the oxidation by Br?2, (SCN)?2 and N3 radicals, respectively. The rate constants do not vary appreciably in alkaline solutions (pH 8.9). The ionic strength dependence was observed for the rate constants of the oxidation by Br?2 and (SCN)?2. Those rate constants estimated on the assumption that the radicals react only with the amino acid residues with the characteristic steric correction factors were less than one-tenth of the observed ones. These results suggest that the partially exposed region of the heme is the probable site of electron transfer from ferrocytochrome c to the radical.Hydroxyl radicals also oxidize ferrocytochrome c with a high rate constant (k > 1 · 1010 M?1 · s?1), but with a very small efficiency (5%).  相似文献   

6.
《Free radical research》2013,47(4):241-253
We have evaluated the abilities of ferulic acid, (±) catechin, (+) catechin and (-) epicatechin to scavenge the reactive oxygen species hydroxyl radical (OH±), hypochlorous acid (HOCl) and peroxyl radicals (RO2).

Ferulic acid tested at concentrations up to 5 mM inhibited the peroxidation of phospholipid liposomes. Both (±) and (+) catechin and (-) epicatechin were much more effective. All the compounds tested reacted with trichloromethyl peroxyl radical (CCl3O2) with rate constants > 1 × 106M?1s?1.

A mixture of FeCl3-EDTA, hydrogen peroxide (H2O2) and ascorbic acid at pH 7.4, has often been used to generate hydroxyl radicals (OH.) which are detected by their ability to cause damage to the sugar deoxyribose. Ferulic acid, (+) and (±) catechin and (-) epicatechin inhibited deoxyribose damage by reacting with OH. with rate constants of 4.5 × 109M?1s?1, 3.65 × 109M?1s?1, 2.36 × 109M?1s?1 and 2.84 × 109M?1s?1 respectively. (-) Epicatechin, ferulic acid and the (+) and (±) catechins exerted pro-oxidant action, accelerating damage to DNA in the presence of a bleomycin-iron complex. On a molar basis, ferulic acid was less effective in causing damage to DNA compared with the catechins.

A mixture of hypoxanthine and xanthine oxidase generates O2 which reduces cytochrome c to ferrocytochrome c. (+) Catechin and (-) epicatechin inhibited the reduction of cytochrome c in a concentration dependent manner. Ferulic acid and (±) catechin had only weak effects.

All the compounds tested were able to scavenge hypochlorous acid at a rate sufficient to protect alpha-1-antiproteinase against inactivation. Our results show that catechins and ferulic acid possess antioxidant properties. This may become important given the current search for “natural” replacements for synthetic antioxidant food additives.  相似文献   

7.
Hao Yu 《Free radical research》2013,47(9-10):1005-1013
Abstract

Pulse radiolysis was conducted to investigate: several fundamental reactions of a natural flavonoid, rutin, and its glycosylated form (αG-rutin) as a basis for their radiation protection properties; the reactions with ?OH (radical scavenging) and dGMP radical, dGMP? (chemical repair), which was used as a model of initial and not yet stabilised damage on DNA. Three absorption peaks were commonly seen in the reactions of the flavonoids with ?OH, showing that their reactive site is the common structure, i.e. aglycone. One among the three peaks was attributed to the flavonoid radical produced as a result of the removal of a hydrogen atom. The same peak was found in their reactions with dGMP?, showing that dGMP? is chemically repaired by obtaining a hydrogen atom supplied from the flavonoids. Such a spectral change due to the chemical repair was as clear as never reported. The rate constants of the chemical repair reaction were estimated as (9?±?2)×108 M?1 s?1 and (6?±?1)×108 M?1 s?1 for rutin and αG-rutin, respectively. The rate constants of the radical scavenging reactions towards ?OH were estimated as (1.3?±?0.3)×1010 M?1 s?1 and (1.0?±?0.1)×1010 M?1 s?1 for rutin and αG-rutin, respectively. In addition, there was no obvious difference between rutin and αG-rutin, indicating that the glycosylation does not change early chemical reactions of rutin.  相似文献   

8.
Modification of tyrosine (TyrOH) is used as a marker of oxidative and nitrosative stress. 3,3′-Dityrosine formation, in particular, reflects oxidative damage and results from the combination of two tyrosyl phenoxyl radicals (TyrO). This reaction is in competition with reductive processes in the cell which ‘repair’ tyrosyl radicals: possible reductants include thiols and ascorbate. In this study, a rate constant of 2 × 106 M−1 s−1 was estimated for the reaction between tyrosyl radicals and glutathione (GSH) at pH 7.15, generating the radicals by pulse radiolysis and monitoring the tyrosyl radical by kinetic spectrophotometry. Earlier measurements have suggested that this ‘repair’ reaction could be an equilibrium, and to investigate this possibility the reduction (electrode) potential of the (TyrO,H+/TyrOH) couple was reinvestigated by observing the fast redox equilibrium with the indicator 2,2′-azinobis(3-ethylbenzothiazoline-6-sulphonate). Extrapolation of the reduction potential of TyrO measured at pH 9–11 indicated the mid-point reduction potential of the tyrosyl radical at pH 7, Em7(TyrO,H+/TyrOH) = 0.93 ± 0.02 V. This is close to the reported reduction potential of the glutathione thiyl radical, Em7 = 0.94 ± 0.03 V, confirming the ‘repair’ equilibrium constant is of the order of unity and suggesting that efficient reduction of TyrO by GSH might require removal of thiyl radicals to move the equilibrium in the direction of repair. Loss of thiyl radicals, facilitating repair of TyrO, can arise either via conjugation of thiyl with thiol/thiolate or oxygen, or unimolecular transformation, the latter important at low concentrations of thiols and oxygen.  相似文献   

9.
Neutral tryptophan (*Trp) and tyrosine (TyrO(*)) radicals are repaired by certain flavonoids in buffer, in micelles and in human serum albumin (HSA) with corresponding formation of semioxidized flavonoid radicals. In deaerated buffer, *Trp but not TyrO(*) radicals react with catechin. In micelles, quercetin and rutin repair both *Trp and TyrO(*) radicals. In addition to amino acid reactivity, microenvironmental factors and nature of the flavonoids govern this repair. Electron transfer efficiencies from quercetin to negatively charged *Trp radicals are 100% in the micellar pseudophases of positively charged cetyltrimethylammonium bromide, (CTAB), and neutral Triton X100 (TX100), but 55% in negatively charged sodium dodecyl sulfate (SDS). In oxygen-saturated CTAB micelles, quercetin also reacts with the superoxide radical anion. When bound to domain IIA of HSA, quercetin repairs, by intra- or intermolecular encounter, less than 20% of oxidative damage to HSA. Quercetin can also repair freely circulating oxidized molecules with repair efficiencies falling to 7% for oxidized Trp, Tyr and alpha-MSH and to less than 2% for urate radical. This limited effectiveness is attributed both to the inaccessibility of bound quercetin and rutin toward radicals of circulating molecules and to the diffusion-controlled recombination of these radicals.  相似文献   

10.
The ‘singlet oxygen scavengers’. 1,4-diazobicyclo(2,2,2)-octane (DABCO), diphenyl furan and β-carotene react rapidly with the organic peroxyradical CCl3O2?. The absolute reaction rate constants k = 1.2 ± 0.2 × 107, 6 ± 2 × 107 at 1.5 ± 0.2 × 109 M?1s?1 respectively have been determined by pulse radiolysis. Comparison with other data suggest that other free radicals are also likely to react with these compounds; in the case of the hydroxyl radical and DABCO k = 1.25 × 109 M?1s?1 has been determined.  相似文献   

11.
12.
Hypochlorite (HOCl/OCl?) and hypobromite (HOBr/OBr?) are shown to react with tert-butyl hydroperoxide with close rate constants (10.8 and 8.9 M?1 s?1, respectively). Using a spin trap α-(4-pyridyl-1-oxide)-N-tert-butyl nitrone, both reactions are shown to proceed through decomposition of the hydroperoxide yielding butylperoxyl [˙OOC(CH3)3] and butoxyl [˙OC(CH3)3] radicals in a ratio depending on the hydroperoxide concentration. Thus, like hypochlorite, hypobromite can generate free radicals in reactions with organic hydroperoxides, which can be important for intensification of free-radical processes, e.g., lipid peroxidation at the chain branching stage.  相似文献   

13.
The interaction of the nonsteroidal anti‐inflammatory drug flurbiprofen (FBP) with human serum albumin (HSA) hardly influences the fluorescence of the protein's single tryptophan (Trp). Therefore, in addition to fluorescence, heavy atom‐induced room‐temperature phosphorescence is used to study the stereoselective binding of FBP enantiomers and their methyl esters to HSA. Maximal HSA phosphorescence intensities were obtained at a KI concentration of 0.2 M. The quenching of the Trp phosphorescence by FBP is mainly dynamic and based on Dexter energy transfer. The Stern–Volmer plots based on the phosphorescence lifetimes indicate that (R)‐FBP causes a stronger Trp quenching than (S)‐FBP. For the methyl esters of FBP, the opposite is observed: (S)‐(FBPMe) quenches more than (R)‐FBPMe. The Stern–Volmer plots of (R)‐FBP and (R)‐FBPMe are similar although their high‐affinity binding sites are different. The methylation of (S)‐FBP causes a large change in its effect on the HSA phosphorescence lifetime. Furthermore, the quenching constants of 3.0 × 107 M?1 s?1 of the R‐enantiomers and 2.5 × 107 M?1 s?1 for the S‐enantiomers are not influenced by the methylation and indicate a stereoselectivity in the accessibility of the HSA Trp to these drugs. Chirality 24:840–846, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

14.
Several rate constants for one-electron reduction of cytochrome P450 are more rapid in the absence than in the presence of the specific substrate. The respective values for methyl viologen, nicotinamide adenine dinucleotide and the 1-methyl-4-(and -3-)carbamidopyridinium radicals are 2.6, 3.4, 6 and 35 × 107 M?1 s?1 without camphor, and 0.15, 0.1, 1.8 and 110 × 107 M?1 s?1 for the camphor complex. Hydrated electrons react with cytochrome P450 with a rate constant of 3.0 × 1010 M?1 s?1 whether camphor is bound or not, but little of the reduction takes place at the haem iron. No reduction of the haem iron by CO2?- or O2?- could be detected, whether camphor is bound or not.  相似文献   

15.
《Free radical research》2013,47(11):1300-1310
Abstract

Hypotaurine and cysteine sulfinic acid are known to be readily oxidized to the respective sulfonates, taurine and cysteic acid, by several oxidative agents that may be present in biological systems. In this work, the relevance of both the carbonate anion and nitrogen dioxide radicals in the oxidation of hypotaurine and cysteine sulfinic acid has been explored by the peroxidase activity of Cu,Zn superoxide dismutase (SOD) and by pulse radiolysis. The extent of sulfinate oxidation induced by the system SOD/H2O2 in the presence of bicarbonate (CO3?– generation), or nitrite (?NO2 generation) has been evaluated. Hypotaurine is efficiently oxidized by the carbonate radical anion generated by the peroxidase activity of Cu,Zn SOD. Pulse radiolysis studies have shown that the carbonate radical anion reacts with hypotaurine more rapidly (k = 1.1 × 109 M?1s?1) than nitrogen dioxide (k = 1.6 × 107 M?1s?1). Regarding cysteine sulfinic acid, it is less reactive with the carbonate radical anion (k = 5.5 × 107 M?1s?1) than hypotaurine. It has also been observed that the one-electron transfer oxidation of both sulfinates by the radicals is accompanied by the generation of transient sulfonyl radicals (RSO2?). Considering that the carbonate radical anion could be formed in vivo at high level from bicarbonate, this radical can be included in the oxidants capable of performing the last metabolic step of taurine biosynthesis. Moreover, the protective effect exerted by hypotaurine and cysteine sulfinate on the carbonate radical anion-mediated tyrosine dimerization indicates that both sulfinates have scavenging activity towards the carbonate radical anion. However, the formation of transient reactive intermediates during sulfinate oxidation by carbonate anion and nitrogen dioxide radical may at the same time promote oxidative reactions.  相似文献   

16.
The protein from Arabidopsis thaliana gene locus At1g79260.1 is comprised of 166‐residues and is of previously unknown function. Initial structural studies by the Center for Eukaryotic Structural Genomics (CESG) suggested that this protein might bind heme, and consequently, the crystal structures of apo and heme‐bound forms were solved to near atomic resolution of 1.32 Å and 1.36 Å, respectively. The rate of hemin loss from the protein was measured to be 3.6 × 10?5 s?1, demonstrating that it binds heme specifically and with high affinity. The protein forms a compact 10‐stranded β‐barrel that is structurally similar to the lipocalins and fatty acid binding proteins (FABPs). One group of lipocalins, the nitrophorins (NP), are heme proteins involved in nitric oxide (NO) transport and show both sequence and structural similarity to the protein from At1g79260.1 and two human homologues, all of which contain a proximal histidine capable of coordinating a heme iron. Rapid‐mixing and laser photolysis techniques were used to determine the rate constants for carbon monoxide (CO) binding to the ferrous form of the protein (k′CO = 0.23 μM?1 s?1, kCO = 0.050 s?1) and NO binding to the ferric form (k′NO = 1.2 μM–1 s–1, kNO = 73 s?1). Based on both structural and functional similarity to the nitrophorins, we have named the protein nitrobindin and hypothesized that it plays a role in NO transport. However, one of the two human homologs of nitrobindin contains a THAP domain, implying a possible role in apoptosis. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

17.
The kinetics of O·-2 reaction with semi-oxidized tryptophan radicals in lysozyme, Trp·(Lyz) have been investigated at various pHs and conformational states by pulse radiolysis. The Trp·(Lyz) radicals were formed by Br·-2 oxidation of the 3–4 exposed Trp residues in the protein. At pH lower than 6.2, the apparent bimolecular rate is about 2 × 108M-1s-1; but drops to 8 × 107M-1s-1 or less above pH 6.3 and in CTAC micelles. Similarly, the apparent bimolecular rate constant for the intermolecular Trp·(Lyz) + Trp·(Lyz) recombination reaction is about (4-7 × 106M-1s-1) at/or below pH 6.2 then drops to 1.3-1.6 × 106M-1s-1 at higher pH or in micelles. This behavior suggests important conformational and/or microenvironmental rearrangement with pH, leading to less accessible semioxidized Trp· residues upon Br·-2 reaction. The kinetics of Trp·(Lyz) with ascorbate, a reducing species rather larger than O·-2 have been measured for comparison. The well-established long range intramolecular electron transfer from Tyr residues to Trp radicals-leading to the repair of the semi-oxidized Trp·(Lyz) and formation of the tyrosyl phenoxyl radical is inhibited by the Trp·(Lyz)+O·-2 reaction, as is most of the Trp·(Lyz)+Trp·(Lyz) reaction. However, the kinetic behavior of Trp·(Lyz) suggests that not all oxidized Trp residues are involved in the intermolecular recombination or reaction with O·-2. As the kinetics are found to be quite pH sensitive, this study demonstrates the effect of the protein conformation on O·-2 reactivity. To our knowledge, this is the first report on the kinetics of a protein-O·-2 reaction not involving the detection of change in the redox state of a prosthetic group to probe the reactivity of the superoxide anion.  相似文献   

18.
Kinetic study of the reaction between tocopheroxyl (vitamin E radical) and reduced ubiquinone, n = 10) has been performed. The rates of reaction of ubiquinol with α-tocopheroxyl 1 and seven kinds of alkyl substituted tocopheroxyl radicals 2–8 in solution have been determined spectrophotometrically, using a stopped-flow technique. The result shows that the rate constants decrease as the total electron-donating capacity of the alkyl substituents on the aromatic ring of tocopheroxyls increases. For the tocopheroxyls with two alkyl substituents at ortho positions (C-5 and C-7), the second-order rate constants, k1, obtained vary i n the order of 102, and decrease predominantly, as the size of two ortho-alkyl groups (methyl, ethyl, isopropyl and tert-buty) in tocopheroxyl increases. On the other hand, the reaction between tocopheroxyl and ubiquinone-10 (oxidized ubiquinone) has not been observed. The result indicates that ubiquinol-10 regenerates tocopherol by donating a hydrogen atom of the 1-OH and/or 4-OH group to the tocopheroxyl radical. For instance, the k1 values obtained for α-tocopheroxyl are 3.74 · 105 M?1 · s?1 and 2.15 · 5 M?1 · s?1 in benzene and ethanol solution at 25°C, respectively. The above reaction rates, k1, obtained were compared with those of vitamin C with α-tocopheroxyl reported by Packer et al. (k2 = 1.55 · 106 M?1 · s?1) and Scarpa et al. (K2 = 2 · 105 105 M?1 · s?1), which is well known as a usual regeneration reaction of tocopheroxyl in biomembrane systems. The result suggests that ubiquinol-10 also regenerates the tocopheroxyl to tocopherol and prevents lipid peroxidation in various tissues and mitochondria.  相似文献   

19.
Helix-coil dynamics of a Z-helix hairpin   总被引:1,自引:0,他引:1  
The helix–coil transition of a Z-helix hairpin formed from d(C-G)5T4(C-G)5 has been characterized by equilibrium melting and temperature jump experiments in 5M NaClO4 and 10 mM Na2HPO4, pH 7.0. The melting curve can be represented by a simple all-or-none transition with a midpoint at 81.6 ± 0.4°C and an enthalpy change of 287 ± 15 kJ/mole. The temperature jump relaxation can be described by single exponentials at a reasonable accuracy. Amplitudes measured as a function of temperature provide equilibrium parameters consistent with those derived from equilibrium melting curves. The rate constants of Z-helix formation are found in the range from 1800 s?1 at 70°C to 800 s?1 at 90°C and are associated with an activation enthalpy of ?(50 ± 10) kJ/mole, whereas the rate constants of helix dissociation are found in the range from 200 s?1 at 70°C to 4500 s?1 at 90°C with an activation enthalpy +235 kJ/mole. These parameters are consistent with a requirement of 3–4 base pairs for helix nucleation. Apparently nucleation occurs in the Z-helix conformation, because a separate slow step corresponding to a B to Z transition has not been observed. In summary, the dynamics of the Z-helix–coil transition is very similar to that of previously investigated right-handed double helices.  相似文献   

20.
Recombination of apo horseradish peroxidase with 2,4 dimethyldeutero hemin and its mono- and dimethyl esters was performed. The number of free carboxyl side chains in these three hemins is 2, 1 and 0 respectively. Despite such a difference, all of these three reconstituted enzymes can react with H2O2 to produce compound I. The second order rate constants for compound I formation are 1.3 × 107 M?1s?1, 8.5 × 106 M?1s?1 and 5.9 × 106 M?1s?1. Therefore the propionate side chain of hemin has no direct role in compound I formation.  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