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.
Phenoxyl radicals generated pulse radiolytically by the reaction of N.3 with Gly-Tyr decay biomolecularly (2k = 4.7 X 10(8)M-1 s-1) with efficient formation of 2,2'-dimers, which enolize rapidly (k = 2.7 X 10(4) s-1) to produce the 2,2'-biphenolic product. The build-up of the characteristic 2,2'-biphenol fluorescence (400 nm) and absorption also indicated a delayed (k = 80 s-1) process, probably involving the phenoxyl <-> phenoxy-quinol equilibrium. About 60 per cent of the Gly-Tyr phenoxyls were found to dimerize to the 2,2'-biphenol, and a similarly efficient 2,2'-coupling seems to occur with other tyrosyls, such as Lys-Tyr-Lys and histone. gamma-Radiolysis was applied to estimate relative yields of formation of 2,2'-biphenols under various conditions. Dimerization is almost completely inhibited by cysteine or oxygen, consistent with phenoxyl 'repair' by cysteine or O-.2; disproportionation of O-.2 with SOD prevents repair. The phenol 2,2'-coupling is less efficient for .OH- and inefficient for e-aq-initiation.  相似文献   

2.
Most of the mechanistic studies of tyrosine nitration have been performed in aqueous solution. However, many protein tyrosine residues shown to be nitrated in vitro and in vivo are associated to nonpolar compartments. In this work, we have used the stable hydrophobic tyrosine analogue N-t-BOC-L-tyrosine tert-butyl ester (BTBE) incorporated into phosphatidylcholine (PC) liposomes to study physicochemical and biochemical factors that control peroxynitrite-dependent tyrosine nitration in phospholipid bilayers. Peroxynitrite leads to maximum 3-nitro-BTBE yields (3%) at pH 7.4. In addition, small amounts of 3,3'-di-BTBE were formed at pH 7.4 (0.02%) which increased over alkaline pH; at pH 6, a hydroxylated derivative of BTBE was identified by HPLC-MS analysis. BTBE nitration yields were similar in dilauroyl- and dimyristoyl-PC and were also significant in the polyunsaturated fatty acid-containing egg PC. *OH and *NO2 scavengers inhibited BTBE nitration. In contrast to tyrosine in the aqueous phase, the presence of CO2 decreased BTBE nitration, indicating that CO3*- cannot permeate to the compartment where BTBE is located. On the other hand, micromolar concentrations of hemin and Mn-tccp strongly enhanced BTBE nitration. Electron spin resonance (ESR) detection of the BTBE phenoxyl radical and kinetic modeling of the pH profiles of BTBE nitration and dimerization were in full agreement with a free radical mechanism of oxidation initiated by ONOOH homolysis in the immediacy of or even inside the bilayer and with a diffusion coefficient of BTBE phenoxyl radical 100 times less than for the aqueous phase tyrosyl radical. BTBE was successfully applied as a hydrophobic probe to study nitration mechanisms and will serve to study factors controlling protein and lipid nitration in biomembranes and lipoproteins.  相似文献   

3.
Nitrogen dioxide is a product of peroxynitrite homolysis and peroxidase-catalyzed oxidation of nitrite. It is of great importance in protein tyrosine nitration because most nitration pathways end with the addition of *NO2 to a one-electron-oxidized tyrosine. The rate constant of this radical addition reaction is high with free tyrosine-derived radicals. However, little is known of tyrosine radicals in proteins. In this paper, we have used *NO2 generated by gamma radiolysis to study the nitration of the R2 subunit of ribonucleotide reductase, which contains a long-lived tyrosyl radical on Tyr122. Most of the nitration occurred on Tyr122, but nonradical tyrosines were also modified. In addition, peptidic bonds close to nitrated Tyr122 could be broken. Nitration at Tyr122 was not observed with a radical-free metR2 protein. The estimated rate constant of the Tyr122 radical reaction with *NO2 was of 3 x 10(4) M(-1) s(-1), thus several orders of magnitude lower than that of a radical on free tyrosine. Nitration rate of other tyrosine residues in R2 was even lower, with an estimated value of 900 M(-1) s(-1). This study shows that protein environment can significantly reduce the reactivity of a tyrosyl radical. In ribonucleotide reductase, the catalytically active radical residue is very efficiently protected against nitrogen oxide attack and subsequent nitration.  相似文献   

4.
E Cadenas  G Merényi  J Lind 《FEBS letters》1989,253(1-2):235-238
The reaction between the phenoxyl radical of Trolox C, a water-soluble vitamin E analogue, and superoxide anion radical was examined by using the pulse radiolysis technique. The results indicate that the Trolox C phenoxyl radical may undergo a rapid one-electron transfer from superoxide radical [k = (4.5 +/- 0.5) x 10(8) M-1.S-1] to its reduced form. This finding indicates that superoxide radical might play a role in the repair of vitamin e phenoxyl radical.  相似文献   

5.
The effect of oxygen on the radiolysis of tyrosine in aqueous solutions was investigated by using gamma and pulsed electron irradiation. Steady-state radiolysis was reexamined and extended to include the effect of pH and determination of hydrogen peroxide. The loss of tyrosine, G(-Tyr), during irradiation and yields of 3,4-dihydroxyphenylalanine, G(DOPA), and hydrogen peroxide, G(H2O2), are determined in the pH range from 1 to 9. In the whole pH range used G(-Tyr) equals G(DOPA), and a higher G(H2O2) than expected was observed. In slightly acid and neutral media, both G(-Tyr) and G(DOPA) equal the yield of hydroxyl radicals, GOH, formed in the radiolysis of water, while the excess of hydrogen peroxide equals 1/2 GOH. Hence it was concluded that all tyrosine OH-adducts react with oxygen yielding peroxy radicals. In acid and alkaline media all measured yields decrease. This is caused by formation of tyrosine phenoxyl radicals (TyrO), which react with superoxide anion (O2-) and hydroperoxy (HO2) radicals regenerating tyrosine. By using pulse radiolysis K(TyrO + O2) less than or equal to 2 X 10(5) mol-1 dm3 s-1 and k(TyrO + O2-) = (1.7 +/- 0.2) X 10(9) mol-1 dm3 s-1 were determined. On the basis of the results, a reaction mechanism is proposed.  相似文献   

6.
A series of radical transitions, Br2-.----Met(S therefore Br)----Trp(indolyl)----Tyr (phenoxyl), has been demonstrated by pulse radiolysis of N2O-saturated aqueous solutions containing Br-, Met-Gly and Trp-(Gly)2-Tyr at pH 6.7. The intramolecular Met(S therefore Br)----Trp(indolyl) transition in the dipeptide Met-Trp is shown to proceed via the Trp+. radical cation, with a rate constant of k approximately 10(7)s-1, consistent with an electron transfer. Br2-.-attack upon ribonuclease A (RNase) leads to a fast Met(S therefore Br)----Tyr(phenoxyl) process, k = (4.0 +/- 1.0) X 10(5)s-1, probably involving the solvent-exposed Met-29 and the adjacent Tyr-25. Phenoxyl dimerization in the RNase system produces the characteristic o,o'-biphenol fluorescence, but a competing interaction of the Tyr-25(phenoxyl) with the 26-84 disulphide group also appears possible.  相似文献   

7.
The iron chelating agent desferrioxamine inhibits peroxynitrite-mediated oxidations and attenuates nitric oxide and oxygen radical-dependent oxidative damage both in vitro and in vivo. The mechanism of protection is independent of iron chelation and has remained elusive over the past decade. Herein, stopped-flow studies revealed that desferrioxamine does not react directly with peroxynitrite. However, addition of peroxynitrite to desferrioxamine in both the absence and the presence of physiological concentrations of CO2 and under excess nitrite led to the formation of a one-electron oxidation product, the desferrioxamine nitroxide radical, consistent with desferrioxamine reacting with the peroxynitrite-derived species carbonate (CO3*-) and nitrogen dioxide (*NO2) radicals. Desferrioxamine inhibited peroxynitrite-dependent free radical-mediated processes, including tyrosine dimerization and nitration, oxyhemoglobin oxidation in the presence of CO2, and peroxynitrite plus carbonate-dependent chemiluminescence. The direct two-electron oxidation of glutathione by peroxynitrite was unaffected by desferrioxamine. The reactions of desferrioxamine with CO3*- and *NO2 were unambiguously confirmed by pulse radiolysis studies, which yielded second-order rate constants of 1.7 x 10(9) and 7.6 x 10(6) M(-1) s(-1), respectively. Desferrioxamine also reacts with tyrosyl radicals with k = 6.3 x 10(6) M(-1) s(-1). However, radical/radical combination reactions between tyrosyl radicals or of tyrosyl radical with *NO2 outcompete the reaction with desferrioxamine and computer-assisted simulations indicate that the inhibition of tyrosine oxidation can be fully explained by scavenging of the peroxynitrite-derived radicals. The results shown herein provide an alternative mechanism to account for some of the biochemical and pharmacological actions of desferrioxamine via reactions with CO3*- and *NO2 radicals.  相似文献   

8.
At pH 7.0 the alkaline mesentericopeptidase has ultraviolet absorption spectrum with a minimum at 251 nm and a maximum at 280 nm and no visible absorption. From the tyrosine to tryptophan ratio a value of 3 tryptophyl residues per mole of protein is obtained. The molar extinction coefficient at 280 nm is 3.55 X 10(4)M-1cm-1. Spectrophotometric titration studies show that the molecule of mesentericopeptidase contains seven phenolic groups with a pKapp - 9.92 and four to five groups with a pKapp = 11.96. Denaturing agents, such as 5 M guanidine hydrochloride or alkali, normalize the ionization of the tyrosyl residues. There is a good correlation between the spectrophotometric titration data and the results for the reactivities of the tyrosines in mesentericopeptidase towards tetranitromethane. The correlation is explained by the mechanism of nitration. Conclusions about the state of the tyrosyl residues and the three-dimensional structure of mesentericopeptidase are made.  相似文献   

9.
The kinetics of reaction of singly reduced methemoglobin (HbFe3(3+)Fe2+) with carbon monoxide have been investigated by the pulse radiolysis method. The rate constant for carbon monoxide binding to this form of hemoglobin is 4.1 X 10(6) M-1 S-1 at 24 degrees in our solutions. This value compares with existing values for various forms of hemoglobin ranging from 4 X 10(6) to 6.5 X 10(6) M-1 S-1. Addition of inositol hexaphosphate to the solutions results in a lower rate constant for carbon monoxide binding amounting to 1.1 X 10(5) M-1 S-1.  相似文献   

10.
Oxidation of tyrosine moieties by radicals involved in lipid peroxidation is of current interest; while a rate constant has been reported for reaction of lipid peroxyl radicals with a tyrosine model, little is known about the reaction between tyrosine and alkoxyl radicals (also intermediates in the lipid peroxidation chain reaction). In this study, the reaction between a model alkoxyl radical, the tert-butoxyl radical and tyrosine was followed using steady-state and pulse radiolysis. Acetone, a product of the β-fragmentation of the tert-butoxyl radical, was measured; the yield was reduced by the presence of tyrosine in a concentration- and pH-dependent manner. From these data, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 6 ± 1 × 10(7) M(-1) s(-1) at pH 10. Tyrosine phenoxyl radicals were also monitored directly by kinetic spectrophotometry following generation of tert-butoxyl radicals by pulse radiolysis of solutions containing tyrosine. From the yield of tyrosyl radicals (measured before they decayed) as a function of tyrosine concentration, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 7 ± 3 × 10(7) M(-1) s(-1) at pH 10 (the reaction was not observable at pH 7). We conclude that reaction involves oxidation of tyrosine phenolate rather than undissociated phenol; since the pK(a) of phenolic hydroxyl dissociation in tyrosine is ≈ 10.3, this infers a much lower rate constant, about 3 × 10(5) M(-1) s(-1), for the reaction between this alkoxyl radical and tyrosine at pH 7.4.  相似文献   

11.
J Everse  N Kujundzic 《Biochemistry》1979,18(12):2668-2673
A detailed investigation of the reduction of cytochrome c by glutathione has shown that the reaction proceeds through several steps. A rapid combination of the reducing agent with the cytochrome leads to the formation of a glutathione-cytochrome intermediate in which the glutathione most likely interacts with the edge of the heme moiety. The electron transfer takes place in a subsequent slower step. Since cytochrome c(III) exists in two conformational forms at neutral pH [Kujundzic, N., & Everse, J. (1978) Biochem. Biophys. Res. Commun. 82, 1211], the reduction of cytochrome c by glutathione may be represented by cyt c(III) + GS- reversible K1 cyt c(III) ... GS- reversible k1 products cyt c*(III) + GS- reversible K2 cyt c*(III) ... GS- reversible k2 products At 25 degrees C, pH 7.5, and an ionic strength of 1.0 (NaCl), k1 = 1.2 X 10(-3) S-1, k2 = 2.0 X 10(-3) S-1, k1 = 2.9 X 10(3) M-1, and K2 = 5.3 X 10(3) M-1. The reaction is catalyzed by trisulfides, and second-order rate constants of 4.55 X 10(3) and 7.14 X 10(3) M-1 S-1 were obtained for methyl trisulfide and cysteine trisulfide, respectively.  相似文献   

12.
The dissociation of nitric oxide from hemoglobin, from isolated subunits of hemoglobin, and from myoglobin has been studied using dithionite to remove free nitric oxide. The reduction of nitric oxide by dithionite has a rate of 1.4 X 10(3) M-1 S-1 at 20 degrees in 0.05 M phosphate, pH 7.0, which is small compared with the rate of recombination of hemoglobin with nitric oxide (25 X 10(6) M-1 S-1 (Cassoly, R., and Gibson, Q. H. (1975) J. Mol. Biol. 91, 301-313). The rate of NO combination with chains and myoglobin was found to be 24 X 10(6) M-1 S-1 and 17 X 10(6) M-1 S-1, respectively. Hence, the observed progress curve of the dissociation of nitric oxide is dependent upon the dithionite concentration and the total heme concentration. Addition of excess carbon monoxide to the dissociation mixture reduces the free heme yielding a single exponential process for chains and for myoglobin which is dithionite and heme concentration independent over a wide range of concentrations. The rates of dissociation of nitric oxide from alpha chains, from beta chains, and from myoglobin are 4.6 X 10(-5) S-1, 2.2 X 10(-5) S-1, and 1.2 X 10(4) S-1, respectively, both in the presence and in the absence of carbon monoxide at 20 degrees in 0.05 M phosphate, pH 7.0. Analogous heme and dithionite concentration dependence is found for the dissociation of nitric oxide from tetrameric hemoglobin. The reaction is cooperative, the intrinsic rate constants for the dissociation of the 1st and 4th molecules of NO differing about 100-fold. With hemoglobin, replacement of NO by CO at neutral pH is biphasic in phosphate buffers. The rate of the slow phase is 1 X 10(-5) S-1 and is independent of pH. The amplitude of the fast phase increases with lowering of pH. By analogy with the treatment of the HbCO + NO reaction given by Salhany et al. (Salhany, J.M., Ogawa, S., and Shulman, R.G. (1975) Biochemistry 14, 2180-2190), the fast phase is attributed to the dissociation of NO from T state molecules and the slow phase to dissociation from R state molecules. Analysis of the data gives a pH-independent value of 0.01 for the allosteric constant c (c = Kr/Kt where Kr and Kt are the dissociation constants for NO from the R and T states, respectively) and pH-dependent values of L (2.5 X 10(7) at pH 7 in 0.05 M phosphate buffer). The value of c is considerably greater than that for O2 and CO. Studies of the difference spectrum induced in the Soret region by inositol hexaphosphate are also reported. This spectrum does not arise directly from the change of conformation between R and T states. The results show that if the equilibrium binding curve for NO could be determined experimentally, it would show cooperativity with Hill's n at 50% saturation of about 1.6.  相似文献   

13.
We have shown previously that peroxynitrite-induced nitration of a hydrophobic tyrosyl probe is greater than that of tyrosine in the aqueous phase (Zhang, H., Joseph, J., Feix, J., Hogg, N., and Kalyanaraman, B. (2001) Biochemistry 40, 7675-7686). In this study, we have tested the hypothesis that the extent of tyrosine nitration depends on the intramembrane location of tyrosyl probes and on the nitrating species. To this end, we have synthesized membrane spanning 23-mer containing a single tyrosyl residue at positions 4, 8, and 12. The location of the tyrosine residues in the phospholipid membrane was determined by fluorescence and electron spin resonance techniques. Nitration was initiated by slow infusion of peroxynitrite, co-generated superoxide and nitric oxide ((.)NO), or a myeloperoxidase/hydrogen peroxide/nitrite anion (MPO/H(2)O(2)/NO(2)(-)) system. Results indicate that with slow infusion of peroxynitrite, nitration of transmembrane tyrosyl peptides was much higher (10-fold or more) than tyrosine nitration in aqueous phase. Peroxynitrite-dependent nitration of tyrosyl-containing peptides increased with increasing depth of the tyrosyl residue in the bilayer. In contrast, MPO/H(2)O(2)/ NO(2)(-)-induced tyrosyl nitration decreased with increasing depth of tyrosyl residues in the membrane. Transmembrane nitrations of tyrosyl-containing peptides induced by both peroxynitrite and MPO/H(2)O(2)/NO(2)(-) were totally inhibited by (.)NO that was slowly released from spermine NONOate. Nitration of peptides in both systems was concentration-dependently inhibited by unsaturated fatty acid. Concomitantly, an increase in lipid oxidation was detected. A mechanism involving (.)NO(2) radical is proposed for peroxynitrite and MPO/H(2)O(2)/NO(2)(-)-dependent transmembrane nitration reactions.  相似文献   

14.
Radiation chemical experiments demonstrate that the reaction of tyrosyl radical (TyrO(.)) with (.)NO(2) yields 45 +/- 3% 3-nitrotyrosine and that a major product of the reaction of TyrO(.) with (.)NO is 3,3'-dityrosine. Radiolysis was used to generate (.)NO and O-(2) in the presence of tyrosine and bicarbonate at pH 7.5 +/- 0.1. The nitration yield was found to be dose rate-dependent, and the yield per radical produced by pulse radiolysis was identical to that obtained with authentic peroxynitrite. The proposed mechanism that accounts for the data is as follows: (i) In the presence of CO(2) the reaction of (.)NO with O-(2) yields 33% (.)NO(2) and CO-(3), where the latter reacts rapidly with tyrosine to form TyrO(.); (ii) The formation of 3-nitrotyrosine takes place via the reaction of (.)NO(2) with TyrO(.), which is the main process at high dose rates; and (iii) Under continuous generation of (.)NO and O-(2), the formation of 3-nitrotyrosine is strongly suppressed because of efficient scavenging of (.)NO(2) by tyrosine. The proposed model shows that the highest nitration yield is obtained for similar fluxes of (.)NO and O-(2) and is completely inhibited upon excess production of O-(2) because of efficient scavenging of TyrO(.) by O-(2). The biological implications of these findings are discussed.  相似文献   

15.
The chain oxidation of lactate dehydrogenase-bound NADH initiated by superoxide radicals and propagated by oxygen was studied with pulse radiolysis. The kinetic parameters were re-evaluated in a system with carefully purified reagents (water and other chemicals) and in the presence of EDTA. The rate constant for the oxidation of the enzyme-bound NADH by O2- is calculated from the observed pseudo-first order disappearance of NADH and the chain length (molecules of NADH oxidized per O2- anion generated in the pulse). It is (1.0 +/- 0.2) X 10(5) M-1 S-1, consistent within a 13-fold variation in lactate dehydrogenase. NADH complex concentration and with varying chain length up to 6.1. Based on experiments with varying pH values from 4.5 to 9.0, the rate constant for oxidation of enzyme-bound NADH by HO2 is estimated to be 2.0 X 10(6) M-1 S-1.  相似文献   

16.
Eosinophil peroxidase (EPO) has been implicated in promoting oxidative tissue injury in conditions ranging from asthma and other allergic inflammatory disorders to cancer and parasitic/helminthic infections. Studies thus far on this unique peroxidase have primarily focused on its unusual substrate preference for bromide (Br(-)) and the pseudohalide thiocyanate (SCN(-)) forming potent hypohalous acids as cytotoxic oxidants. However, the ability of EPO to generate reactive nitrogen species has not yet been reported. We now demonstrate that EPO readily uses nitrite (NO(2)(-)), a major end-product of nitric oxide ((.)NO) metabolism, as substrate to generate a reactive intermediate that nitrates protein tyrosyl residues in high yield. EPO-catalyzed nitration of tyrosine occurred more readily than bromination at neutral pH, plasma levels of halides, and pathophysiologically relevant concentrations of NO(2)(-). Furthermore, EPO was significantly more effective than MPO at promoting tyrosine nitration in the presence of plasma levels of halides. Whereas recent studies suggest that MPO can also promote protein nitration through indirect oxidation of NO(2)(-) with HOCl, we found no evidence that EPO can indirectly mediate protein nitration by a similar reaction between HOBr and NO(2)(-). EPO-dependent nitration of tyrosine was modulated over a physiologically relevant range of SCN(-) concentrations and was accompanied by formation of tyrosyl radical addition products (e.g. o,o'-dityrosine, pulcherosine, trityrosine). The potential role of specific antioxidants and nucleophilic scavengers on yields of tyrosine nitration and bromination by EPO are examined. Thus, EPO may contribute to nitrotyrosine formation in inflammatory conditions characterized by recruitment and activation of eosinophils.  相似文献   

17.
A kinetic study of the reaction between vitamin C (L-ascorbic acid, AsH2) and a tocopheroxyl radical (7-tert-butyl-5-isopropyltocopheroxyl) in Triton X-100 micellar solution has been performed using stopped-flow spectrophotometry. The second-order rate constants (k2) obtained showed notable pH dependence with a broad maximum around pH 8. For instance, the k2 values obtained were 26 M-1 S-1 at pH 3, 322 M-1 S-1 at pH 7, and 273 M-1 S-1 at pH 10. A good correlation between the rate constants and the mole fraction of ascorbate monoanion (AsH-) was observed, showing that ascorbate (AsH-) can regenerate the tocopherol from tocopheroxyl in biological systems. Furthermore, the results indicate that reduced ascorbic acid (AsH2) does not have the ability to regenerate the tocopherol in aqueous solution. On the other hand, it was found that AsH2 can reduce the tocopheroxyl to tocopherol in benzene/ethanol (2:1) mixtures, although the rate of reaction is only approximately 15% of that observed in micellar solution at pH 7.  相似文献   

18.
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. At 20 degrees C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4-10(6) 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 approximately 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 = 1.4-10(6) M-1 - S-1, the form above pH 7.45 with k = 3.0- 10(5) M-1 - S-1, and the form present above pH 9.2 with k = 0. 3. The reaction has an activation energy of 20 kJ mol-1 and an enthalpy of activation at 25 degrees 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. 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 5-10(5)-5-10(6) M-1 - S-1.  相似文献   

19.
1-[2-Amino-5-(6-carboxyindol-2-yl)phenoxyl]-2-(2'- amino-5'-methylphenoxy)ethane-N,N,N',N'-tetraacetic acid (indo-1) and 2-[2-(bis(carboxymethyl)amino-5-methylphenoxy) methyl]-6- methyl-8-[bis-(carboxymethyl)amino]quinoline (quin-2) are sensitive, spectral indicators for Zn2+. Additions of subsaturating Zn2+ to 10-80 microM indo-1 or quin-2 at pH 7.0 produce uv difference spectra with isosbestic wavelengths at 342 and 282 nm or at 342, 317, and 252 nm, respectively. Formation of 1:1 Zn2+:indicator complexes at pH 7.0 and 20 degrees C in the absence (presence) of 100 mM KCl gives delta epsilon max = -2.4 +/- 0.2 X 10(4) M-1 cm-1 at 367 nm (-2.1 +/- 0.2 X 10(4) M-1 cm-1 at 365 nm) for indo-1 and delta epsilon max = -2.7 +/- 0.1 X 10(4) M-1 cm-1 at 266 nm (-2.6 +/- 0.1 X 10(4) M-1 cm-1 at 265 nm) for quin-2. Competition experiments at pH 7.0 and 20 degrees C with indo-1 and quin-2 and also 4-(2-pyridylazo)resorcinol (PAR) as the second chelator in the absence (presence) of 100 mM KCl yield apparent affinity constants: K'A = 2.5 +/- 1.0 X 10(10) M-1 (6.2 +/- 0.5 X 10(9) M-1) for indo-1 binding Zn2+ and K'A = 9.4 +/- 3.3 X 10(11) M-1 (2.7 +/- 0.1 X 10(11) M-1) for quin-2 binding Zn2+. The above constants provide the basis for rapid steady-state spectrophotometric determinations of the affinity of a protein for Zn2+ with K'A approximately 10(10) - 10(13) M-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Transferrin and the transferrin model compound Fe(III)-EHPG (Fe(III)-ethylenediamine-N,N'-bis[2-(2-hydroxyphenyl)acetic acid] were found not to react with superoxide, as pulse radiolysis and kinetic spectroscopy revealed no transient species and no bleaching of the 465-nm absorption. However, transferrin was found to react with the formate radical, CO-.2, and the hydrated electron, e-aq, with second-order rate constants of 3.8 X 10(8) and 1.1 X 10(10) M-1 S-1, respectively. These reactions produced a transient species (lambda max = 420 nm) which subsequently decayed by a second-order process. However, no reduction of the Fe(III) in transferrin was detected. Fe(III)-EHPG was also found to react with CO-.2 and e-aq, k = 7.3 X 10(6) and 1.1 X 10(9) M-1 S-1, respectively. The reactions of CO-.2 and e-aq with Fe(III)-EHPG resulted in no transient species but rather in reduction of the iron. These results are consistent with the inability of transferrin and Fe(III)-EHPG to catalyze the Haber-Weiss reaction.  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