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 46 毫秒
1.
Using conductivity detection, pulse radiolysis experiments showed that solvent protonation of the electron adducts of cytosine, 5-methyl cytosine and 2'-deoxycytidine occurs with rate constants k greater than or equal to 2 x 10(4) M-1S-1. The protonated electron adducts transfer an electron to p-nitroactetophenone (PNAP) with rate constants ranging from 3.5 x 10(9) to 5.3 x 10(9) M-1S-1. The transfer is quantitative (G = 2.7), as shown by conductometric and spectroscopic measurements. In the presence of O2 no electron transfer to O2 takes place, implying that O2 adds to the protonated electron adduct radicals. No electron transfer from the H- and OH-adducts of the cytosine derivatives, either to PNAP or to O2, takes place near neutral pH. It is suggested that the differences in the reaction behaviour of the H-adduct radicals and the protonated electron adduct radicals towards PNAP can be accounted for if different radicals are formed by H-addition and protonation of the electron adduct. The H atoms most probably add to the C-5-C-6 double bonds, whereas the electron adducts are protonated at N-3 and/or 0-2.  相似文献   

2.
The putative intermediate dienol (2) in the steroid isomerase (KSI) catalyzed conversion of 5-androstene-3,17-dione (1) to 4-androstene-3,17-dione (3) has been independently generated and tested as a substrate for KSI. At pH 7, dienol 2 is converted by KSI to a mixture of 1 (46%) and 3 (54%). The apparent second-order rate constant for reaction of 2 with KSI to produce 3 (kappa cat/Km = 2.3 x 10(8) M-1 s-1) is similar to that for reaction of 1 with KSI (kappa cat/Km = 2.1 x 10(8) M-1 s-1), demonstrating that 2 is kinetically competent. Isomerization of 1 by KSI in D2O gives only 5% of solvent deuterium incorporated into the product 3. When 2 reacts with KSI in D2O, and the product 3 is isolated (from direct reaction of 2 and from subsequent conversion of the 1 initially formed), ca. 80 atom % deuterium is located at C-6 beta, confirming that protonation of the dienol by KSI occurs at the same face as the proton transfer in the KSI catalyzed reaction of 1 to 3.  相似文献   

3.
《Free radical research》2013,47(6):393-405
Hahn-Meitner-Institut Berlin, Bereich Strahlenchemie, Glienicker Str. 100, 1000 Berlin 39, Fed. Rep. Germany

Penicillamine thiyl radicals, PenS', are shown to abstract hydrogen atoms from 2-propanol and to establish an equilibrium

PenS' + (CH1)2 CHOH ? PenSH + (CH3)2 COH.

The rate constants for the forward and back reaction have been determined to (1.4 ± 0.3) × 104 and (1.2 ± 0.3) × 108 M-1 s-1, respectively, by means of pulse radiolysis. The data have been obtained from various independent methods which include direct measurements and competitive schemes involving irreversible interception of the alcohol radical by electron acceptors (e.g. CCl4. PNAP) and/or the thiyl radical by antioxidants (e.g. α-tocopherol). The results demonstrate that the reaction of carbon-centered radicals with thiols, in radiation biology commonly known as “repair” reaction, may be reversed and thus imply the possibility of thiyl radical induced biological damage.  相似文献   

4.
One-electron oxidation of (6R)-5,6,7,8-tetrahydrobiopterin (H(4)B) by the azide radical generates the radical cation (H(4)B(*)(+)) which rapidly deprotonates at physiological pH to give the neutral trihydrobiopterin radical (H(3)B(*)); pK(a) (H(4)B(*)(+) <==> H(3)B(*) + H(+)) = (5.2 +/- 0.1). In the absence of ascorbate both the H(4)B(*)(+) and H(3)B(*) radicals undergo disproportionation to form quinonoid dihydrobiopterin (qH(2)B) and the parent H(4)B with rate constants k(H(4)B(*)(+) + H(4)B(*)(+)) = 6.5 x 10(3) M(-1) s(-1) and k(H(3)B(*) + H(3)B(*)) = 9.3 x 10(4) M(-1) s(-1), respectively. The H(3)B(*) radical is scavenged by ascorbate (AscH(-)) with an estimated rate constant of k(H(3)B(*) + AscH(-)) similar 1.7 x 10(5) M(-1) s(-1). At physiological pH the pterin rapidly scavenges a range of biological oxidants often associated with cellular oxidative stress and nitric oxide synthase (NOS) dysfunction including hydroxyl ((*)OH), nitrogen dioxide (NO(2)(*)), glutathione thiyl (GS(*)), and carbonate (CO(3)(*-)) radicals. Without exception these radicals react appreciably faster with H(4)B than with AscH(-) with k(*OH + H(4)B) = 8.8 x 10(9) M(-1) s(-1), k(NO(2)(*) + H(4)B) = 9.4 x 10(8) M(-1) s(-1), k(CO(3)(*-) + H(4)B) = 4.6 x 10(9) M(-1) s(-1), and k(GS(*) + H(4)B) = 1.1 x 10(9) M(-1) s(-1), respectively. The glutathione disulfide radical anion (GSSG(*-)) rapidly reduces the pterin to the tetrahydrobiopterin radical anion (H(4)B(*-)) with a rate constant of k(GSSG(*-) + H(4)B) similar 4.5 x 10(8) M(-1) s(-1). The results are discussed in the context of the general antioxidant properties of the pterin and the redox role played by H(4)B in NOS catalysis.  相似文献   

5.
Sulfite radical anion, SO3-., which is generated either by non-enzymatic reaction of hydrogen peroxide (H2O2-) with sulfite (SO3(2-)) or by the oxidation of bisulfite (HSO3) with Ce4+ ion, can be trapped with a water-soluble, nitroso-aromatic spin-trap, sodium 3,5-dibromo-4-nitrosobenzenesulfonate (DBNBS, 1), yielding an ESR spectrum with coupling constants [aN (1) = 12.9 G, aH (2) = 0.8 G] and a g-value of 2.0063. The SO3- radical adduct (spin adduct) was observed even in the presence of the very low concentration of H2O2 (1.21 X 10(-2) mumol).  相似文献   

6.
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   

7.
The rate constant for the reduction of the tyrosyl radical with selenocysteine has been measured to investigate whether selenocysteine is capable of repair of protein radicals. Tyrosyl radicals, both free in solution and in insulin, were generated by means of pulse radiolysis and laser flash photolysis in aqueous solution. The rate constant for the reaction of free N-acetyl-tyrosyl-amine radicals with selenocysteine is (8 +/- 2) x 10 (8) M (-1) s (-1), and that for tyrosyl radicals in insulin is (1.6 +/- 0.4) x 10 (8) M (-1) s (-1). The rate constant for the reaction of selenoglutathione with the N-acetyl-tyrosyl-amine radical is (5 +/- 2) x 10 (8) M (-1) s (-1). In contrast, cysteine and glutathione react more slowly than their selenium analogues with the tyrosyl radical: the reactions of N-acetyl-tyrosyl-amine radicals with cysteine and glutathione are 3 and 5 orders of magnitude slower, respectively, than those with selenocysteine and selenoglutathione, while those of tyrosyl radicals in insulin are 3 and 2 orders of magnitude slower, respectively.  相似文献   

8.
Kinetic study of the reaction between vitamin E radical and alkyl hydroperoxides has been performed, as a model for the reactions of lipid hydroperoxides with vitamin E radical in biological systems. The rates of reaction of hydroperoxides (n-butyl hydroperoxide 1, sec-butyl hydroperoxide 2, and tert-butyl hydroperoxide 3) with vitamin E radical (5,7-diisopropyl-tocopheroxyl 4) in benzene solution have been determined spectrophotometrically. The second-order rate constants, k-1, obtained are 1.34 x 10(-1) M-1s-1 for 1, 2.42 x 10(-1) M-1s-1 for 2, and 3.65 x 10(-1) M-1s-1 for 3 at 25.0 degrees C. The result indicates that the rate constants increase as the total electron donating capacity of the alkyl substituents at alpha-carbon atom of hydroperoxides increases. The above rates, k-1, are about seven order of magnitude lower than those, k1, for the reaction of vitamin E with peroxyl radical.  相似文献   

9.
The reactions of Fe(II)EDTA, Fe(II)DTPA, and Fe(II)HEDTA with hydrogen peroxide near neutral pH have been investigated. All these reactions have been assumed to proceed through an active intermediate, I1, (Formula: see text) where pac is one of the three polyaminocarboxylates mentioned above. I1, whether .OH radical or an iron complex, reacts with ethanol, formate, and other scavengers at rates relative to k2 that, with the exception of t-butanol and benzoate, are similar, but not identical, to those expected for the.OH radical. In contrast, at pH 3, in the absence of ligands the reaction of I1 with Fe2+ was inhibited by ethanol and t-butanol and the reactivity of I1 towards these two scavengers relative to ferrous ion is identical to that exhibited by the hydroxyl radical. When pac = HEDTA, the intermediate of the first reaction reacts with formate ion to form the ferrous HEDTA ligand radical complex, which is characterized by absorption maxima at 295 nm (epsilon = 2,640 M-1 cm-1) and 420 nm (epsilon = 620 M-1 cm-1). For the reaction of Fe(II)HEDTA with H2O2, the following mechanism is proposed: (Formula: see text) where k17 = 4.2 X 10(4) M-1 sec-1 and k19 = 5 +/- 0.2 sec-1.  相似文献   

10.
N,N-Dimethyl-p-phenylenediamine (DMPD) reacted directly with oxyhemoglobin under formation of ferrihemoglobin and, presumably, the N,N-dimethyl-p-phenylenediamine radical cation (DMPP.+). The apparent second-order rate constant of this reaction was 1 M-1 s-1 (pH 7.4, 37 degrees C). The reaction rate was diminished by catalase (by 1/3) and by superoxide dismutase (by 1/5). The apparent second-order rate constant of ferrihemoglobin formation by DMPD.+ was 5 x 10(3) M-1 s-1. Since DMPD.+ is disproportionated by 50% at pH 7.4, the quinonediimine could not be excluded as the ultimate ferrihemoglobin forming oxidant. To prove this hypothesis, the disproportionation equilibrium was shifted to the radical side by addition of excess DMPD. Ferrihemoglobin formation was thereby increased, indication that the radical was the responsible oxidant. In contrast to ferrihemoglobin formation, reactions with glutathione occurred predominantly with the quinonediimine. The second-order rate constant of this reaction was 4 x 10(5) M-1 s-1 which approaches the value obtained with p-benzoquinone. In contrast to the corresponding reactions of the N,N,N',N'-tetramethyl-p-phenylenediamine radical cation, the disporportionation reaction of DMPD.+ was very fast, k = 2 x 10(6) M-1 s-1. Formation of glutathione disulfide was negligible and the main reaction products were two isomeric glutathione adducts, 2- and 3-(glutathione-S-yl)-N,N-dimethyl-p-phenylenediamine. In human erythrocytes, DMPD produced many equivalents of ferrihemoglobin, diminished glutathione and produced both thioethers. In contrast to ferrihemoglobin formation, DMPD and glutathione disappearance as well as thioether appearance occured only after a marked lag phase. The calculated steady state concentration of DMPD.+ was only 4 x 10(-6) the DMPD concentration, as long as ferrihemoglobin was low. At increasing ferrihemoglobin higher steady state concentrations of the radical are attained. In fact, preformed ferrihemoglobin in red cells significantly accelerated DMPD and glutathione disappearance. This effect was completely prevented in the presence of ferrihemoglobin-complexing cyanide. The presented experiments once more appoint blood as a metabolically competent organ for the biotransformation of aromatic amines.  相似文献   

11.
The one-electron reduction product of 1-methyl-4-phenyl-2,3-dihydropyridinium ion has been generated by pulse radiolysis and its absorption spectrum recorded. This radical was found to decay by second-order kinetics (2k = 9.5 x 10(8) M-1 s-1) to 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine and 1-methyl-4-phenyl-2,3-dihydropyridinium ion. Reactions of the above radical species and that formed by one-electron reduction of 1-methyl-4-phenylpyridinium ion, which can also be generated by one-electron oxidation of 1-methyl-4-phenyl-1,2-dihydropyridine, with a number of molecules of biochemical interest have been studied. The one-electron reduction product of oxidised nicotinamide adenine dinucleotide efficiently reduced 1-methyl-4-phenyl-2,3-dihydropyridinium ion (k = 2.2 x 10(9) M-1 s-1). The relevance of these results in relation to redox cycling, a possible mechanism for 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine neurotoxicity, is discussed.  相似文献   

12.
Ion formation from the reaction of triplet (T) and ground state (P) octaethyl-porphyrin (OEP) and zinc octaethyl porphyrin (ZnOEP) and the corresponding cross-reactions have been measured in dry acetonitrile. A uniquely sensitive and fast conductance apparatus and a pulsed dye laser allowed the measurements to be made at the necessarily very low concentrations of T. The hemogeneous reaction of T (ZnOEP) and P (ZnOEP) occurs with rat constant k(1) = 2.0 x 10(8) M(-1)s(-1) and an ion yield of 67%. The similar homogeneous reaction of OEP has k(2) = 1.3 x 10(8)M(-1)s(-1) but an ion yield of only 3%. The cross-reaction of T (OEP) with P (ZnOEP) has k(3) = 1.5 x 10(8) M(-1)s(-1) and an ion yield of 27%, while the inverse cross-reaction of T (ZnOEP) with P (OEP) has k(4) = 3 x 10(8) M(-1)s(-1) and an ion yield of 20%. Thus, the rate constants are only slightly affected but the yields are sensitive to the porphyrin. The possible formation of the heterogeneous ions ZnOEP+ + OEP-, thermodynamically favored by 0.3 V over the homogeneous ions, has little influence on the observed yields. The data are explained by electron transfer and Coulomb field-electon spin-controlled escape of the initial ion-pair.  相似文献   

13.
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   

14.
The sulphate radical SO4(.-) reacts with 1,3-dimethyluracil (1,3-DMU) (k = 5 X 10(9) dm3 mol-1 s-1) thereby forming with greater than or equal to 90 per cent yield the 1,3-DMU C(5)-OH adduct radical 4 as evidenced by its absorption spectrum and its reactivity toward tetranitromethane. Pulse-conductometric experiments have shown that a 1,3-DMU-SO4(.-) aduct 3 as well as the 1,3-DMU radical cation 1, if formed, must be very short-lived (t1/2 less than or equal to 1 microsecond). The 1,3-DMU C(5)-OH adduct 4 reacts slowly with peroxodisulphate (k = 2.1 X 10(5) dm3 mol-1 s-1). It is suggested that the observed new species is the 1,3-DMU-5-OH-6-SO4(.-) radical 7. At low dose rates a chain reaction is observed. The product of this chain reaction is the cis-5,6-dihydro-5,6-dihydroxy-1,3-dimethyluracil 2. At a dose rate of 2.8 X 10(-3) Gys-1 a G value of approximately 200 was observed ([1,3-DMU] = 5 X 10(-3) mol dm-3; [S2O8(2-)] = 10(-2) mol dm-3; [t-butanol] = 10(-2) mol dm-3). The peculiarities of this chain reaction (strong effect of [1,3-DMU], smaller effect of [S2O(2-)8]) is explained by 7 being an important chain carrier. It is proposed that 7 reacts with 1,3-DMU by electron transfer, albeit more slowly (k approximately 1.2 X 10(4) dm3 mol-1 s-1) than does SO4(.-). The resulting sulphate 6 is considered to hydrolyse into 2 and sulphuric acid which is formed in amounts equivalent to those of 2. Computer simulations provide support for the proposed mechanism. The results of some SCF calculations on the electron distribution in the radical cations derived from uracil and 1-methyluracil are also presented.  相似文献   

15.
A mechanism for beta-chlorophenethylamine inhibition of dopamine beta-monooxygenase has been postulated in which bound alpha-aminoacetophenone is generated followed by an intramolecular redox reaction to yield a ketone-derived radical cation as the inhibitory species (Mangold, J.B., and Klinman, J.P. (1984) J. Biol. Chem. 259, 7772-7779). Based on the assumption that the ketone radical is the inhibitory intermediate, an analogous system was predicted and verified (Bossard, M.J., and Klinman, J.P. (1986) J. Biol. Chem. 261, 16421-16427). In the present study, the role of alpha-aminoacetophenone as the proposed intermediate in the inactivation by beta-chlorophenethylamine was examined in greater detail. From the interdependence of tyramine and alpha-aminoacetophenone concentrations, ketone inactivation is concluded to occur at the substrate site as opposed to potential binding at the reductant-binding site. Using beta-[2-1H]- and beta-[2-2H]chlorophenethylamine, the magnitude of the deuterium isotope effect on inactivation under second-order conditions has been found to be identical to that observed under catalytic turnover, D(kappa inact/Ki) = D(kappa cat/Km) = 6-7. By contrast, the isotope effect on inactivation under conditions of substrate and oxygen saturation, D kappa inact = 2, is 3-fold smaller than that seen on catalytic turnover, D kappa cat = 6. This reduced isotope effect for inactivation is attributed to a normal isotope effect on substrate hydroxylation followed by an inverse isotope effect on the partitioning of the enol of alpha-aminoacetophenone between oxidation to a radical cation versus protonation to regenerate ketone. These findings are unusual in that two isotopically sensitive steps are present in the inactivation pathway whereas only one is observable in turnover.  相似文献   

16.
The relaxation rates of the carbon-bound protons and of the three assigned phosphorus resonances of propionyl-CoA were measured in solutions of free propionyl-CoA and of the transcarboxylase-propionyl-CoA complex. In free propionyl-CoA, analysis of the 1/T1 values of 15 protons at 100 and 220 MHz and of 1/T1 and 1/T2 of the three phosphorus atoms at 40.5 MHz indicated free rotation of the propionyl region (taur approximately 3 x 10(-11) sec) but hindered motion of the remainder of the molecule with correlation times of 1-3. 5 x 10(-10) sec, approaching the tumbling time of the entire molecule (taur - 6 x 10(-10) sec. The correlation times of the three phosphorus atoms were indistinguishable from those of their nearest neighbor protons. The effects of three homogeneous enzyme preparations with varying contents of Zn(II), Co(II), and Cu(II) on 1/T1 of 12 protons and 3 phosphorus atoms of prionyl-CoA were analyzed with the help of simultaneous equations to yield the individual contributions at the three metal sites. Only diamagnetic effects were detected on the relaxation rates of the three phosphorus atoms. From the diamagnetic effects it was calculated that the motions of the prionyl side chain and of the terminal pantetheine methylene protons were hindered on the enzyme by an order of magnitude (taur approximately 6 x 10(-10) sec) and that the phosphorus atoms were hindered by two orders of magnitude (taur approximately 1 x 10(-8) sec) over the taur values found in free propionyl-CoA, but that these taur values remained well below that of the entire protein molecule (taur =6 x 10(-7) sec)...  相似文献   

17.
The dynamic behavior of various types of cytochromes c in the redox reaction with iron hexacyanides was studied using a temperature-jump method in order to elucidate the molecular mechanism of the redox reaction of cytochromes with their oxidoreductants. Transmittance after the temperature jump changed through a single exponential decay for all cytochromes investigated. Under a constant concentration of anion, the redox reaction of various types of cytochrome c with iron hexacyanides was analyzed according to the scheme: (see formula in text) where C(III) and C(II) are ferric and ferrous cytochromes, respectively, Fe(III) and Fe(II) are ferri- and ferrocyanides, respectively, C(III) . Fe(II) is the ferricytochrome-ferrocyanide complex and C(II) . Fe(III) is the ferrocytochrome-ferricyanide complex. When step B is slower than the other two steps A and C, tau-1 can be represented approximately as (see formula in text) where the bar over the variables denotes the equilibrium value. In a large excess of ferrocyanide against cytochrome, we can estimate kappa 2, kappa-2, K1 and K3 independently. In the case of horse cytochrome c at 18 degrees C in 0.1 M phosphate buffer at pH 7 with 0.3 M KNO3, the estimated parameters are kappa 2 = 100 +/- 50 S-1, kappa-2 = (3.5 +/- 1.0) . 10(3) S-1, K1 = 15 +/- 7 M-1 and K3 = (8.5 +/- 1.5). 10(-4) M. From the same experiments for seven cytochromes (cytochrome c from horse, tuna, Candida krusei, Saccharomyces oviformis, Rhodospirillum rubrum cytochrome c2, Spirulina platensis cytochrome c-554 and Thermus thermophilus cytochrome c-552), the following results can be deduced. (1) Each parameter defined in the scheme above (kappa 2, kappa-2, K1, K3) diverged beyond the error range. Above all, kappa 2 values of cytochromes c-554 and c-552 are as large as 1 . 10(4) S-1 and much larger than those for the other cytochromes (to 50 approx. 700 S-1). (2) The variance of kappa 2K1 and kappa-2/K3 are relatively less than the variances of individual parameters (kappa 2, kappa-2, K1 and K3), which suggests that the values of kappa 2K1 and kappa-2/K3 have been conserved during the course of evolution.  相似文献   

18.
The processes of aggregation and refolding of recombinant human creatine kinase (rHCK) were studied. Most of the rHCK expressed in E. coli was present in the insoluble traction and it could be solubilized in 6 M urea solution. Unfolding of rHCK in 6 M urea showed biphasic kinetic courses (kappa1 = 6.5 x 10(-3) s(-1); kappa2 = 0.54 x 10(-3) s(-1)) as observed by maximum fluorescence wavelength change. During refolding of the rHCK dissolved in urea, significant aggregation was noticed following first-order kinetics. Aggregation rate constants were influenced by the concentration of NaCl, which increased the difference in transition-free energy (deltadeltaG), showing that stabilization of folding intermediates by NaCl could efficiently reduce the formation of insoluble aggregates. Formations of aggregate were also reduced by adjusting temperature, pH, and concentration of rHCK. Refolding of rHCK under the optimized condition which prevented the aggregation also showed multi-kinetic phases (kappa1 = 3.0 x 10(-3) s(-1); kappa2 = 0.64 x 10(-3) s(-1)). Under optimized conditions applied in this study, rHCK could correctly refold retrieving the high specific enzymatic activity.  相似文献   

19.
Absolute rate-constants for the reaction of the nitroxyl free radicals TAN and TMPN with radiation-chemically-formed radicals and ions have been determined. k(TAN + X) (in M(-1) sec(-1)=4-0 X 10(9) (for X = OH-), 2-9 X 10(10) (eaq-), 8-0 X 10(9) (H-), 7-2 X 10(8) (-CH2OH), 4-0 X 10(8) (CH3CHOH), 4-3 X 10(8) ((CH3)2COH) 2-8 X 10(8) (-CH2(CH3)2COH), 5-9 X 10(7) (glucose radical), 4-0 X 10(8) (c-C5H9-), and k(TMPN + X)=3-4 X 10(9) (OH-), 7-8 X 10(9) (eq-), 4-9 X 10(9) (H-), 4-4 X 10(8) (-CH2OH), 4-9 X 10(8) (CH3CHOH), 3-6 X 10(8) ((CH3)2COH), 1-5 X 10(8) (-CH2(CH3)2COH), 4-9 X 10(7) (glucose radical), 4-3 X 10(8) (c-C5H9-). Direct measurements by means of a pulse-radiolysis conductivity technique were based on the formation and destruction of charged species in these reactions within certain pH ranges. It is indicated that the radiosensitizing nitroxyls undergo both redox and addition reactions.  相似文献   

20.
We discovered that 4-(2-pyridylazo) resorcinol (PAR) has a strong catalytic effect on luminol-potassium ferricyanide chemiluminescence (CL). Results indicated that the chemiluminescence intensities at maximum light emission were linearly corrected with the concentration of PAR over the range 1.0 x 10(-5)-1.0 x 10(-7) mol/L. A detection limit of 5.7 x 10(-8) mol/L for PAR was achieved. It was found that some metal ions strongly affected this catalytic reaction. Based on this finding, the luminol-potassium ferricyanide-PAR reaction was developed for the determination of metal ions. The detection limits (S/N = 3) for Ni2+, Cr3+, Zn2+, Co2+ and Mn2+ were determined to be 1.0 x 10(-9) mol/L, 5.0 x 10(-9) mol/L, 5.0 x 10(-8) mol/L, 1.0 x 10(-9) mol/L and 1.0 x 10(-8) mol/L, respectively. In addition, the relative standard deviation values for these metal ion assays were in the range 0.82-2.72% (n = 6).  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