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.
Getoff N 《Radiation research》2000,154(6):692-696
The spectroscopic and kinetic characteristics of beta-carotene radical cation (beta-carotene(.+)) were studied by pulse radiolysis in aerated DMSO solution. The buildup of beta-carotene(.+) with k(1) = (4.8 +/- 0.2) x 10(8) dm(3) mol(-1) s(-1) [lambda(max) = 942 nm, epsilon = (1.6 +/- 0.1) x 10(4) dm(3) mol(-1) cm(-1)] results from an electron transfer from beta-carotene to DMSO(.+). The beta-carotene(.+) species decays exclusively by first-order reaction, k = (2.1 +/- 0.1) x 10(3) s(-1), probably by two processes: (1) at low substrate concentration by hydrolysis and (2) at high concentrations also by formation of dimer radical cation (beta-carotene)(2)(.+). Under the experimental conditions, a small additional beta-carotene triplet-state absorption ((3)beta-carotene) in the range of 525 to 660 nm was observed. This triplet absorption is quenched by oxygen (k = 7 x 10(4) s(-1)), resulting in singlet oxygen ((1)O(2)), whose reactions can also lead to additional formation of beta-carotene(.+).  相似文献   

2.
Pan J  Lin W  Wang W  Han Z  Lu C  Yao S  Lin N  Zhu D 《Biophysical chemistry》2001,89(2-3):193-199
By use of pulse radiolysis techniques, the radical cations of purine nucleotides have been successfully produced by the SO4- ion oxidation. Time-resolved spectroscopic evidence is provided that the one-electron-oxidized radicals of dAMP and dGMP can be efficiently repaired by aromatic amino acids (including tyrosine and tryptophan) via electron transfer reaction. As a model peptide, Arg-Tyr-AcOH was also investigated with regard to its interaction with deprotonated purine radical cations. The rate constants of the electron transfer reactions were determined to be (1 approximately 5) x 10(8) dm(3) mol(-1) s(-1). These results suggest that the aromatic amino acids in DNA-associated proteins may play some role in electron transfer reactions through DNA.  相似文献   

3.
Shi Y  Kang J  Lin W  Fan P  Jia Z  Yao S  Wang W  Zheng R 《Biochimica et biophysica acta》1999,1472(1-2):279-289
The repair effects on deoxynucleotide radical cations of phenylpropanoid glycosides (PPGs) and their analogs, isolated from a Chinese folk medicinal herb, were studied using the pulse radiolysis technique. The radical cations of deoxynucleotides were formed by the reaction of SO4*- with deoxynucleotides. On pulse irradiation of a nitrogen saturated deoxynucleotide aqueous solution containing 20 mM K2S2O8, 200 mM t-BuOH and one of the PPGs or their analogs, the transient absorption spectra of the radical cations of nucleotide decayed with the formation of those of the radical cation of PPGs or their analogs within several tens of microseconds after electron pulse irradiation. The result indicates that deoxynucleotide radical cations can be repaired by PPGs or their analogs. The rate constants of the repair reactions were determined to be 0.48-1.1 x 10(9), 0.64-1.80 x 10(9) and 2.12-4.4 x 10(9) M(-1) s(-1) for dAMP, dGMP and dCMP radical cations respectively. It is obvious that the rate constants of the repair reaction depend on the number of phenolic hydroxyl groups contained in the PPGs and their analogs. A deeper understanding of this new repair mechanism will undoubtedly help researchers design strategies to prevent and/or intervene more effective in free radical related diseases.  相似文献   

4.
Bacteria can reduce toxic and carcinogenic Cr(VI) to insoluble and less toxic Cr(III). Thermus scotoductus SA-01, a South African gold mine isolate, has been shown to be able to reduce a variety of metals, including Cr(VI). Here we report the purification to homogeneity and characterization of a novel chromate reductase. The oxidoreductase is a homodimeric protein, with a monomer molecular mass of approximately 36 kDa, containing a noncovalently bound flavin mononucleotide cofactor. The chromate reductase is optimally active at a pH of 6.3 and at 65 degrees C and requires Ca(2+) or Mg(2+) for activity. Enzyme activity was also dependent on NADH or NADPH, with a preference for NADPH, coupling the oxidation of approximately 2 and 1.5 mol NAD(P)H to the reduction of 1 mol Cr(VI) under aerobic and anaerobic conditions, respectively. The K(m) values for Cr(VI) reduction were 3.5 and 8.4 microM for utilizing NADH and NADPH as electron donors, respectively, with corresponding V(max) values of 6.2 and 16.0 micromol min(-1) mg(-1). The catalytic efficiency (k(cat)/K(m)) of chromate reduction was 1.14 x 10(6) M(-1) s(-1), which was >50-fold more efficient than that of the quinone reductases and >180-fold more efficient than that of the nitroreductases able to reduce Cr(VI). The chromate reductase was identified to be encoded by an open reading frame of 1,050 bp, encoding a single protein of 38 kDa under the regulation of an Escherichia coli sigma(70)-like promoter. Sequence analysis shows the chromate reductase to be related to the old yellow enzyme family, in particular the xenobiotic reductases involved in the oxidative stress response.  相似文献   

5.
Using pulse radiolysis and laser flash photolysis, we have investigated the reactions of the deleterious species, e(-)(aq), HO&z.rad;, O(2)(*)(-) and O(2)((1)Delta(g)) with 10 water-soluble cyclopropyl-fused C(60) derivatives including a mono-adduct dendro[60]fullerene (d) and C(60) derivatives based on C(60)[C(COOH)(2)](n=2-6), some of which are known to be neuroprotective in vivo. The rate constants for reactions of e(-)(aq) and HO&z.rad; lie in the range 0.5-3.3 x 10(10) M(-1) s(-1). The d and bis-adduct monoanion radicals display sharp absorption peaks around 1000 nm (epsilon = 7 000-11 500 M(-1) cm(-1)); the anions of the tris-, tetra-, and penta-adduct derivatives have broader, weaker absorptions. The monohydroxylated radicals have their most intense absorption maxima around 390-440 nm (epsilon = 1000-3000 M(-1) cm(-1)). The anion and hydroxylated radical absorption spectra display a blue-shift as the number of addends increases. The radical anions react with oxygen (k approximately 10(7)-10(9) M(-1) s(-1)). The reaction of O(2)(*)(-) with the C(60) derivatives does not occur via an electron transfer. The rate constants for singlet oxygen reaction with the dendrofullerene and eee-derivative in D(2)O at pH 7.4 are k approximately 7 x 10(7) and approximately 2 x 10(7) M(-1) s(-1) respectively, in contrast to approximately 1.2 x 10(5) M(-1) s(-1) for the reaction with C(60) in C(6)D(6). The large acceleration of the rates for electron reduction and singlet oxygen reactions in water is due to a solvophobic process.  相似文献   

6.
Electron transfer within rat neuronal nitric-oxide synthase (nNOS) was investigated by pulse radiolysis. Radiolytically generated 1-methyl-3-carbamoyl pyridinium (MCP) radical was found to react predominantly with the heme of the enzyme with a second-order rate constant for heme reduction of 3 x 10(8) m(-1) s(-1). In the calmodulin (CaM)-bound enzyme a subsequent first-order phase was observed which had a rate constant of 1.2 x 10(3) s(-1). In the absence of CaM, this phase was absent. Kinetic difference spectra for nNOS reduction indicated that the second phase consisted of heme reoxidation accompanied by formation of a neutral flavin semiquinone, suggesting that it is heme to flavin electron transfer. Experiments with the heme proximal surface mutant, K423E, had no second phase, confirming that the mutation blocks interdomain electron transfer. With the autoinhibitory loop deletion mutant, Delta40, the slow phase was observed even in the absence of CaM consistent with the role of the loop in impeding interdomain electron transfer. The rate of heme to FMN electron transfer observed in the wild-type enzyme is approximately 1000 times faster than the FMN to heme electron transfer rate predicted during catalysis from kinetic modeling, suggesting that the catalytic process is slowed by kinetic gating.  相似文献   

7.
Cytochrome c oxidase catalyzes the one-electron oxidation of four molecules of cytochrome c and the four-electron reduction of dioxygen to water. The process involves a number of intramolecular electron-transfer reactions, one of which takes place between the two hemes of the enzyme, hemes a and a3, with a rate of approximately 3 x 10(5) s(-1) (tau approximately 3 micros). In a recent report [Verkhovsky et al. (2001) Biochim. Biophys. Acta 1506, 143-146], it was suggested that the 3 x 10(5) s(-1) electron transfer may be controlled by structural rearrangements and that there is an additional electron transfer that is several orders of magnitude faster. In the present study, we have reinvestigated the spectral changes occurring in the nanosecond and microsecond time frames after photolysis of CO from the fully reduced and mixed-valence enzymes. On the basis of the differences between them, we conclude that in the bovine enzyme the microscopic forward and reverse rate constants for the electron-transfer reactions from heme a to heme a3 are not faster than approximately 2 x 10(5) and approximately 1 x 10(5) s(-1), respectively.  相似文献   

8.
The rapid and reversible electron transfer reaction of cytochrome b562 was observed at an In2O3 electrode. The estimated heterogeneous electron transfer rate constant (k0') was k0' > or = 5.0 x 10(-3) cm s(-1) at pH 6.5. When the methionine-7 (Met-7) residue, which coordinates to the heme iron as an axial ligand, of the wild-type cytochrome b562 was replaced by an Ala or Gly residue, a water molecule bound to the heme iron and the electron transfer rate constants decreased to 1.3 x 10(-3) and 1.8 x 10(-3) cm s(-1), respectively. This decrease in the electron transfer rate would be due to the larger reorganization energy for the structural change at the redox site. The midpoint potential of cytochrome b562 was shifted negatively by approximately 135 mV by replacing Met-7 with Ala or Gly. Similar dissociation kinetics of cyanide for the mutated molecules as compared to native myoglobin was obtained.  相似文献   

9.
M A Cusanovich  G Tollin 《Biochemistry》1980,19(14):3343-3347
Cytochrome c-552 from Chromatium vinosum is an unusual heme protein in that it contains two hemes and one flavin per molecule. To investigate whether intramolecular electron transfer occurs in this protein, we have studied its reduction by external photoreduced flavin by using pulsed-laser excitation. This approach allows us to measure reduction kinetics on the mirosecond time scale. Both fully reduced lumiflavin and lumiflavin semiquinone radical reduce cytochrome c-552 with second-order rate constants of approximately 1.4 x 10(6) M-1s-1 and 1.9 x 10(8) M-1 s-1, respectively. Kinetic and spectral data and the results of similar studies with riboflavin indicate that both the flavin and heme moieties of cytochrome c-552 are reduced simultaneously on a millisecond time scale, with the transient formation of a protein-bound flavin anion radical. This is suggested to be due to rapid intramolecular electron transfer. Further, steric restrictions play an important role in the reduction reaction. Studies were conducted on the redox processes following photolysis of CO-ferrocytochrome c-552 in which the flavin was partly oxidized to resolve the kinetics of electron transfer between the heme and flavin of cytochrome c-552. Based on these results, we conclude that intramolecular electron transfer from ferrous heme to oxidized flavin occurs with a first-order rate constant of greater than 1.4 x 10(6) s-1.  相似文献   

10.
A comparative study using laser flash photolysis of the kinetics of reduction and intramolecular electron transfer among the redox centers of chicken liver xanthine dehydrogenase and of bovine milk xanthine oxidase is described. The photogenerated reductant, 5-deazariboflavin semiquinone, reacts with the dehydrogenase (presumably at the Mo center) in a second-order manner, with a rate constant (k = 6 x 10(7) M-1 s-1) similar to that observed with the oxidase [k = 3 x 10(7) M-1 s-1; Bhattacharyya et al. (1983) Biochemistry 22, 5270-5279]. In the case of the dehydrogenase, neutral FAD radical formation is found to occur by intramolecular electron transfer (kobs = 1600 s-1), presumably from the Mo center, whereas with the oxidase the flavin radical forms via a bimolecular process involving direct reduction by the deazaflavin semiquinone (k = 2 x 10(8) M-1 s-1). Biphasic rates of Fe/S center reduction are observed with both enzymes, which are due to intramolecular electron transfer (kobs approximately 100 s-1 and kobs = 8-11 s-1). Intramolecular oxidation of the FAD radical in each enzyme occurs with a rate constant comparable to that of the rapid phase of Fe/S center reduction. The methylviologen radical, generated by the reaction of the oxidized viologen with 5-deazariboflavin semiquinone, reacts with both the dehydrogenase and the oxidase in a second-order manner (k = 7 x 10(5) M-1 s-1 and 4 x 10(6) M-1 s-1, respectively). Alkylation of the FAD centers results in substantial alterations in the kinetics of the reaction of the viologen radical with the oxidase but not with the dehydrogenase. These results suggest that the viologen radical reacts directly with the FAD center in the oxidase but not in the dehydrogenase, as is the case with the deazaflavin radical. The data support the conclusion that the environments of the FAD centers differ in the two enzymes, which is in accord with other studies addressing this problem from a different perspective [Massey et al. (1989) J. Biol. Chem. 264, 10567-10573]. In contrast, the rate constants for intramolecular electron transfer among the Mo, FAD, and Fe/S centers in the two enzymes (where they can be determined) are quite similar.  相似文献   

11.
The time course of electron transfer in vitro between soluble domains of the Rieske iron-sulfur protein (ISP) and cytochrome f subunits of the cytochrome b(6)f complex of oxygenic photosynthesis was measured by stopped-flow mixing. The domains were derived from Chlamydomonas reinhardtii and expressed in Escherichia coli. The expressed 142-residue soluble ISP apoprotein was reconstituted with the [2Fe-2S] cluster. The second-order rate constant, k(2)((ISP-f)) = 1.5 x 10(6) m(-1) s(-1), for ISP to cytochrome f electron transfer was <10(-2) of the rate constant at low ionic strength, k(2)((f-PC))(> 200 x 10(6) m(-1) s(-1)), for the reduction of plastocyanin by cytochrome f, and approximately 1/30 of k(2)((f-PC)) at the ionic strength estimated for the thylakoid interior. In contrast to k(2)((f-PC)), k(2)((ISP-f)) was independent of pH and ionic strength, implying no significant role of electrostatic interactions. Effective pK values of 6.2 and 8.3, respectively, of oxidized and reduced ISP were derived from the pH dependence of the amplitude of cytochrome f reduction. The first-order rate constant, k(1)((ISP-f)), predicted from k(2)((ISP-f)) is approximately 10 and approximately 150 times smaller than the millisecond and microsecond phases of cytochrome f reduction observed in vivo. It is proposed that in the absence of electrostatic guidance, a productive docking geometry for fast electron transfer is imposed by the guided trajectory of the ISP extrinsic domain. The requirement of a specific electrically neutral docking configuration for ISP electron transfer is consistent with structure data for the related cytochrome bc(1) complex.  相似文献   

12.
Aqueous solution of 2'-deoxyguanosine (5 X 10(-4) M, pH 7.0) was irradiated with 60Co gamma-rays under O2, N2, N2O, and t-BuOH-N2, respectively. A marked increase in fluorescence emission intensity was observed under all atmospheric conditions as was observed in aqueous solutions of adenine and 2'-deoxyadenosine. However, the fluorescence yield from 2'-deoxyguanosine with radiation was lower under O2 and much higher under t-BuOH-N2 than that from 2'-deoxyadenosine though it was not so different both under N2 and N2O. Such high fluorescence yield from 2'-deoxyguanosine especially under t-BuOH-N2 suggests that guanine base has a specific reactivity with hydrated electron or t-butanol radical differing from the other nucleobases.  相似文献   

13.
Arrhenius parameters for the reactions of oxidizing hydroxyl radicals and reducing hydrated electrons with cisplatin, transplatin and carboplatin in aqueous solution have been determined using pulsed electron radiolysis and absorption spectroscopy techniques. Under physiological pH and chloride concentration conditions, hydroxyl radical reaction rate constants of (9.99 +/- 0.20) x 10(9), (8.38 +/- 0.55) x 10(9), and (6.03 +/- 0.08) x 10(9) M(-1) s(-1) at 24.0, 20.7 and 24.0 degrees C, respectively, with corresponding activation energies of 12.79 +/- 0.57, 13.88 +/- 1.14, and 14.35 +/- 0.56 kJ mol(-1) for these three reactions, were determined. These oxidations of cisplatin and transplatin to form a Pt(III) transient are significantly faster than reported previously at room temperature. The lower rate constant for carboplatin is consistent with hydroxyl radical reaction partitioning between reaction at the platinum center and the cyclobutanedicarboxylate ligand. The equivalent reductive hydrated electron reaction rate constants measured were (1.99 +/- 0.04) x 10(10) (24.0 degrees C), (1.77 +/- 0.08) x 10(10) (22.0 degrees C), and (8.92 +/- 0.06) x 10(9) M(-1) s(-1) (24.0 degrees C), with corresponding activation energies of 15.75 +/- 1.00, 19.74 +/- 1.82, and 19.99 +/- 0.34 kJ mol(-1). Again, the values determined for cisplatin and transplatin are faster than reported; however, all three values are consistent with direct reduction of the platinum center to form a Pt(I) moiety.  相似文献   

14.
The photoreduction of 1,4-benzoquinone (BQ), 1,4-naphthoquinone (NQ), 9,10-anthraquinone (AQ) and several derivatives, e.g. dimethylBQ, trimethylBQ, duroquinone, bromoNQ, methoxyNQ, methylAQ and dimethylAQ in acetonitrile-water by ascorbate was studied by time-resolved UV-vis spectroscopy using 20 ns laser pulses at 308 nm and continuous 254 nm irradiation. The semiquinone radical (*QH/Q*(-)) is formed after H-atom transfer from ascorbate to the quinone triplet state. The rate constant for quenching is k(q)=(2-9) x 10(9) M(-1) s(-1). Termination of the radicals takes place in the micros-ms range. The results are compared with those initiated by electron transfer from DABCO under similar conditions, where the k(q) values are similar, but the termination of Q*(-) takes place by electron back transfer not yielding hydroquinones. Specific properties of the quinone triplet state, e.g. self-quenching, nucleophilic water addition and the effects of structure are discussed.  相似文献   

15.
In this investigation of radical formation and reaction in gamma- irradiated DNA and model compounds, we report the conversion of the guanine cation radical (one-electron oxidized guanine, G(.+)) to the C1' sugar radical and another sugar radical at the C3' or C4' position (designated C3'(.)/C4'(.)) by visible and UV photolysis. Electron spin resonance (ESR) spectroscopic investigations were performed on salmon testes DNA as well as 5'-dGMP, 3'-dGMP, 2'-deoxyguanosine and other nucleosides/nucleotides as model systems. DNA samples (25- 150 mg/ml D(2)O) were prepared with Tl(3+) or Fe(CN)(3-)(6) as electron scavengers. Upon gamma irradiation of such samples at 77 K, the electron-gain path in the DNA is strongly suppressed and predominantly G(.+) is found; after UV or visible photolysis, the fraction of the C1' sugar radical increases with a concomitant reduction in the fraction of G(.+). In model systems, 3'- dGMP(+.) and 5'-dGMP(+.) were produced by attack of Cl(.-)(2) on the parent nucleotide in 7 M LiCl glass. Subsequent visible photolysis of the 3'-dGMP(+.) (77 K) results predominantly in formation of C1'(.) whereas photolysis of 5'-dGMP(+.) results predominantly in formation of C3'(.)/C4'(.). We propose that sugar radical formation is a result of delocalization of the hole in the electronically excited base cation radical into the sugar ring, followed by deprotonation at specific sites on the sugar.  相似文献   

16.
Li W  Zheng R  Jia Z  Zou Z  Lin N 《Biophysical chemistry》1997,67(1-3):281-286
Repair effects on thymine radical anion by six phenylpropanoid glycosides (PPGs), isolated from Pedicularis species, were studied using pulse radiolysis method. The thymine radical anion was produced by the reaction of hydrated electron with thymine. PPGs were added into the thymine solution saturated with N(2). Kinetic analysis showed that transient absorption spectrum of thymine radical anion formed at first, and then after several microseconds of pulse radiolysis changed to that of PPG radical anion. The evidence indicated that thymine radical anion was repaired through one-electron transfer between the radical anion and PPG. Electrophilic phenyl-substituted unsaturated carboxylic group containing in PPGs' structure was able to capture electron from thymine radical anion before it undergo reversible protonation. The reaction rate constants of electron transfer from thymine radical anion to PPGs were within 1.16-2.29 x 10(9) dm(3) mol(-1) s(-1).  相似文献   

17.
Time-resolved in situ radiolysis ESR (electron spin resonance, equivalently EPR, electron paramagnetic resonance) studies have shown that the scavenging of radiolytically produced hydroxyl radical in nitrous oxide-saturated aqueous solutions containing 2 mM DMPO is essentially quantitative (94% of the theoretical yield) at 100 micros after the electron pulse [1]. This result appeared to conflict with earlier results using continuous cobalt-60 gamma radiolysis and hydrogen peroxide photolysis, where factors of 35 and 33% were obtained, respectively [2,3]. To investigate this discrepancy, nitrogen-saturated aqueous solutions containing 15 mM DMPO were cobalt-60 gamma irradiated (dose rate = 223 Gy/min) for periods of 0.25-6 min, and ESR absorption spectra were observed approximately 30 s after irradiation. A rapid, pseudo-first-order termination reaction of the protonated DMPO-hydrated electron adduct (DMPO-H) with DMPO-OH was observed for the first time. The rate constant for the reaction of DMPO-H with DMPO-OH is 2.44 x 10(2) (+/- 2.2 x 10(1)) M(-1) s(-1). In low-dose radiolysis experiments, this reaction lowers the observed yield of DMPO-OH to 44% of the radiation-chemical OH radical yield (G = 2.8), in good agreement with the earlier results [2,3]. In the absence of the DMPO-H radical, the DMPO-OH exhibits second-order radical termination kinetics, 2k(T) = 22 (+/- 2) M(-1) s(-1) at initial DMPO-OH concentrations > or = 13 microM, with first-order termination kinetics observed at lower concentrations, in agreement with earlier literature reports [4].  相似文献   

18.
Chromium(VI) salts are well known to be mutagens and carcinogens and to easily cross the cell membranes. Because they are powerful oxidizing agents, Cr(VI) reacts with intracellular materials to reduce to trivalent form, which binds DNA. This study was designed to investigate the interaction of calf thymus DNA with Cr(VI) and Cr(III) in aqueous solution at pH 6.5-7.5, using Cr(VI)/DNA(P) molar ratios (r) of 1:20 to 2:1 and Cr(III)/DNA(P) molar ratios (r) of 1:80 to 1:2. UV-visible and Fourier transform infrared (FTIR) difference spectroscopic methods were used to determine the metal ion-binding sites, binding constants, and the effect of cation complexation on DNA secondary structure. Spectroscopic results showed no interaction of Cr(VI) with DNA at low anion concentrations (r = 1:20 to 1:1), whereas some perturbations of DNA bases and backbone phosphate were observed at very high Cr(VI) contents (r > 1) with overall binding constant of K = 508 M(-1). Cr(III) chelates DNA via guanine N-7 and the nearest PO(2) group with overall binding constant of K = 3.15 x 10(3) M(-1). Evidence for cation chelate formation comes from major shiftings and intensity variations of the guanine band at 1717 and the phosphate asymmetric stretching vibration at 1222 cm(-1). At low Cr(III) concentration (r = 1:40), the number of Cr(III) ions bound to DNA were 6-7 cations/500 base pairs, and this increased to 30-35 cations/500 base pairs at high metal ion content (r = 1:4). DNA condensation occurred at high cation concentration (r = 1:10). No major alteration of DNA conformation was observed, and the biopolymer remained in the B family structure upon chromium complexation.  相似文献   

19.
We directly measured the absolute reactivity of 17-beta-estradiol (E2) and several phenolic model compounds for E2 toward t-butoxy radical (t-BuO*) by nanosecond time-resolved optical spectroscopy. Compared to other phenols, E2 is a moderate, but not strong deactivator of oxyradicals. The absolute bimolecular rate constant for H-atom transfer from E2 to t-BuO* is 1.3 +/- 0.3 x 10(9) M(-1) x s(-1) (23 degrees C, benzene). We estimate the O-H bond strength of 17-beta-estradiol to be approximately 85 +/- 2 kcal/mol and calculate the reaction rate constant of E2 toward peroxy radical to be 10(5) M(-1) x s(-1) at 37 degrees C. The conjugate phenoxy radical of 17-beta-estradiol, E2O*, is unusually reactive toward alpha-tocopherol and ascorbate by H-atom transfer in homogeneous solution (10(8)-10(9) M(-1) x s(-1)). Our findings suggest that E2 functions in vivo as a highly localized, synergistic biological antioxidant. This may partly explain the clinical effectiveness of ovarian steroids in delaying the manifestations of Alzheimer's Disease as well as in protecting against cardiovascular pathologies. In the absence of complementary antioxidant synergists, E2O* is expected to be a pro-oxidant.  相似文献   

20.
A method for the quantitation of the superoxide radical generation rate (V) in murine liver nuclei by the oxidation of 1-hydroxy-2,2,6,6-tetramethyl-4-oxo-piperidine O2-. radicals with the formation of a stable nitroxyl radical recorded by the EPR method, has been developed. It was shown that NADP- and NADPH-dependent superoxide radical generation is suppressed by superoxide dismutase (approximately by 90%). The Km values for NADH and NADPH are 1.5 x 10(-6) and 4.4 x 10(-7) M, respectively; the maximal rate (0.2 nmol.min-1.mg protein-1) is equal for both substrates. Cyanide (greater than 2 mM) causes a practically complete inhibition of the O2-. generation by both substrates. It is suggested that there exists a single readily autooxidized site of O2-. generation by both substrates for NADH- and NADPH-dependent site of the electron transport chain in nuclear membranes.  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