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.
Toxicities of the radiolytically generated oxidizing radicals HO(*), CO(3)(-)(*), and NO(2)(*) toward suspension cultures of a bacterium (Escherichia coli) and a yeast (Saccharomyces cerevisiae) were examined. As demonstrated by the absence of protection from the membrane-impermeable HO(*) scavenger polyethylene glycol (PEG), externally generated HO(*) was not bactericidal under these conditions; however, partial protection by PEG was observed for S. cerevisiae, indicating the presence of a fungicidal pathway involving external HO(*). For both organisms, conversion of external HO(*) to the secondary radical, CO(3)(-)(*), by reaction with HCO(3)(-) increases their susceptibility to radiolytic killing. In contrast, externally generated NO(2)(*) exhibited toxicity comparable to that of CO(3)(-)(*) toward E. coli, but completely blocked the extracellular toxicity of HO(*) toward S. cerevisiae. Cogeneration of equal fluxes of NO(2)(-)(*) and either HO(*) or CO(3)(-)(*) also essentially eliminated the extracellular microbicidal reactions. This behavior is consistent with expectations based upon relative rates of radical-radical self-coupling and cross-coupling reactions. The different patterns of toxicity observed imply fundamentally different microbicidal mechanisms for the two organisms, wherein the bacterium is susceptible to killing by oxidation of highly reactive targets on its cellular envelope but, despite undergoing similar oxidative insult, the fungus is not.  相似文献   

2.
Boccini F  Herold S 《Biochemistry》2004,43(51):16393-16404
The strong oxidizing and nitrating agent peroxynitrite has been shown to diffuse into erythrocytes and oxidize oxyhemoglobin (oxyHb) to metHb. Because the value of the second-order rate constant for this reaction is on the order of 10(4) M(-)(1) s(-)(1) and the oxyHb concentration is about 20 mM (expressed per heme), this process is rather fast and oxyHb is considered a sink for peroxynitrite. In this work, we showed that the reaction of oxyHb with peroxynitrite, both in the presence and absence of CO(2), proceeds via the formation of oxoiron(iv)hemoglobin (ferrylHb), which in a second step is reduced to metHb and nitrate by its reaction with NO(2)(*). In the presence of physiological relevant amounts of CO(2), ferrylHb is generated by the reaction of NO(2)(*) with the coordinated superoxide of oxyHb (HbFe(III)O(2)(*)(-)). This reaction proceeds via formation of a peroxynitrato-metHb complex (HbFe(III)OONO(2)), which decomposes to generate the one-electron oxidized form of ferrylHb, the oxoiron(iv) form of hemoglobin with a radical localized on the globin. CO(3)(*)(-), the second radical formed from the reaction of peroxynitrite with CO(2), is also scavenged efficiently by oxyHb, in a reaction that finally leads to metHb production. Taken together, our results indicate that oxyHb not only scavenges peroxynitrite but also the radicals produced by its decomposition.  相似文献   

3.
Peroxynitrite (ONOO(((-)))/ONOOH) is expected in vivo to react predominantly with CO(2), thereby yielding NO(2)(.) and CO(3) radicals. We studied the inhibitory effects of ascorbate on both NADH and dihydrorhodamine 123 (DHR) oxidation by peroxynitrite generated in situ from 3-morpholinosydnonimine N-ethylcarbamide (SIN-1). SIN-1 (150 micrometer)-mediated oxidation of NADH (200 micrometer) was half-maximally inhibited by low ascorbate concentrations (61-75 micrometer), both in the absence and presence of CO(2). Control experiments performed with thiols indicated both the very high antioxidative efficiency of ascorbate and that in the presence of CO(2) in situ-generated peroxynitrite exclusively oxidized NADH via the CO(3) radical. This fact is attributed to the formation of peroxynitrate (O(2)NOO(-)/O(2)NOOH) from reaction of NO(2)(.) with O(2), which is formed from reaction of CO(3) with NADH. SIN-1 (25 micrometer)-derived oxidation of DHR was half-maximally inhibited by surprisingly low ascorbate concentrations (6-7 micrometer), irrespective of the presence of CO(2). Control experiments performed with authentic peroxynitrite revealed that ascorbate was in regard to both thiols and selenocompounds much more effective to protect DHR. The present results demonstrate that ascorbate is highly effective to counteract the oxidizing properties of peroxynitrite in the absence and presence of CO(2) by both terminating CO(3)/HO( small middle dot) reactions and by its repair function. Ascorbate is therefore expected to act intracellulary as a major peroxynitrite antagonist. In addition, a novel, ascorbate-independent protection pathway exists: scavenging of NO(2)(.) by O(2) to yield O(2)NOO(-), which further decomposes into NO(2)(-) and O(2).  相似文献   

4.
Heo J  Campbell SL 《Biochemistry》2006,45(7):2200-2210
Ras GTPases cycle between inactive GDP-bound and active GTP-bound states to modulate a diverse array of processes involved in cellular growth control. We have previously shown that both NO/O(2) (via nitrogen dioxide, (*)NO(2)) and superoxide radical anion (O(2)(*)(-)) promote Ras guanine nucleotide dissociation. We now show that hydrogen peroxide in the presence of transition metals (i.e., H(2)O(2)/transition metals) and peroxynitrite also trigger radical-based Ras guanine nucleotide dissociation. The primary redox-active reaction species derived from H(2)O(2)/transition metals and peroxynitrite is O(2)(*)(-) and (*)NO(2), respectively. A small fraction of hydroxyl radical (OH(*)) is also present in both. We also show that both carbonate radical (CO(3)(*)(-)) and (*)NO(2), derived from the mixture of peroxynitrite and bicarbonate, facilitate Ras guanine nucleotide dissociation. We further demonstrate that NO/O(2) and O(2)(*)(-) promote Ras GDP exchange with GTP in the presence of a radical-quenching agent, ascorbate, or NO, and generation of Ras-GTP promotes high-affinity binding of the Ras-binding domain of Raf-1, a downstream effector of Ras. S-Nitrosylated Ras (Ras-SNO) can be formed when NO serves as a radical-quenching agent, and hydroxyl radical but not (*)NO(2) or O(2)(*)(-) can further react with Ras-SNO to modulate Ras activity in vitro. However, given the lack of redox specificity associated with the high redox potential of OH(*), it is unclear whether this reaction occurs under physiological conditions.  相似文献   

5.
M Cudic  C Ducrocq 《Nitric oxide》2000,4(2):147-156
To investigate the protective effect of the anesthetic 2, 6-diisopropylphenol, or propofol, in oxidative processes in which (*)NO and peroxynitrite are involved, direct interactions were explored. The reactions of the highly lipophilic propofol with (*)NO in methanolic or aqueous buffered solutions under air were shown to produce the same compounds as those detected with peroxynitrite, but with very low yields and slow rates. In aqueous neutral medium, peroxynitrite (ONOO(-), ONOOCO(-)(2), ONOOH) was able to nitrate and oxidize propofol: In addition to oxidation products, quinone and quinone dimer, the formation of the 4-nitropropofol derivative was detected, increasing with peroxynitrite or CO(2) concentrations. Nitration reached 20% after the addition of 25 mM bicarbonate to an equimolecular mixture of peroxynitrite and propofol in methanol/phosphate-buffered solution (1/4,v/v) at pH 7.4. However, peroxynitrite either in methanol or in alkaline-buffered mixture (optimum pH 10-12) resulted in the rapid and almost complete transformation of propofol to an intermediate compound 1, which further decomposed to 4-nitrosopropofol. The transient compound 1 was obtained from either peroxynitrite or (*)NO in the presence of oxygen. From mass spectrometry determination of compound 1 we propose the involvement of the nitrosodioxyl radical ONOO(*), forming an adduct with the propofoxyl radical, to yield 4-nitrosodioxypropofol and finally 4-nitrosopropofol.  相似文献   

6.
The photodetachment of NO from [M(II)(CN)5NO]2- with M = Fe, Ru, and Os, upon laser excitation at various wavelengths (355, 420, and 480 nm) was followed by various techniques. The three complexes showed a wavelength-dependent quantum yield of NO production Phi(NO), as measured with an NO-sensitive electrode, the highest values corresponding to the larger photon energies. For the same excitation wavelength the decrease of Phi(NO) at 20 degrees C in the order Fe > Ru > Os, is explained by the increasing M-N bond strength and inertness of the heavier metals. Transient absorption data at 420 nm indicate the formation of the [M(III)(CN)5H2O]2- species in less than ca. 1 micros for M = Fe and Ru. The enthalpy content of [Fe(III)(CN)5H2O]2- with respect to the parent [Fe(II)(CN)5NO]2- state is (190 +/- 20) kJ mol(-1), as measured by laser-induced optoacoustic spectroscopy (LIOAS) upon excitation at 480 nm. The production of [Fe(III)(CN)5H2O]2- is concomitant with an expansion of (8 +/- 3) ml mol(-1) consistent with an expansion of the water bound through hydrogen bonds to the CN ligands plus the difference between NO release into the bulk and water entrance into the first coordination sphere. The activated process, as indicated by the relatively strong temperature dependence of the Phi(NO) values and by the temperature dependence of the appearance of the [Fe(III)(CN)5H2O]2- species, as determined by LIOAS, is attributed to NO detachment in less than ca. 100 ns from the isonitrosyl (ON) ligand (MS1 state).  相似文献   

7.
The effect of Zn(2+) or Cu(2+) ions on Mn-depleted photosystem II (PS II) has been investigated using EPR spectroscopy. In Zn(2+)-treated and Cu(2+)-treated PS II, chemical reduction with sodium dithionite gives rise to a signal attributed to the plastosemiquinone, Q(A)(*)(-), the usual interaction with the non-heme iron being lost. The signal was identified by Q-band EPR spectroscopy which partially resolves the typical g-anisotropy of the semiquinone anion radical. Illumination at 200 K of the unreduced samples gives rise to a single organic free radical in Cu(2+)-treated PS II, and this is assigned to a monomeric chlorophyll cation radical, Chl a(*)(+), based on its (1)H-ENDOR spectrum. The Zn(2+)-treated PS II under the same conditions gives rise to two radical signals present in equal amounts and attributed to the Chl a(*)(+) and the Q(A)(*)(-) formed by light-induced charge separation. When the Cu(2+)-treated PS II is reduced by sodium ascorbate, at >/=77 K electron donation eliminates the donor-side radical leaving the Q(A)(*)(-) EPR signal. The data are explained as follows: (1) Cu(2+) and Zn(2+) have similar effects on PS II (although higher concentrations of Zn(2+) are required) causing the displacement of the non-heme Fe(2+). (2) In both cases chlorophyll is the electron donor at 200 K. It is proposed that the lack of a light-induced Q(A)(*)(-) signal in the unreduced Cu(2+)-treated sample is due to Cu(2+) acting as an electron acceptor from Q(A)(*)(-) at low temperature, forming the Cu(+) state and leaving the electron donor radical Chl a(*)(+) detectable by EPR. (3) The Cu(2+) in PS II is chemically reducible by ascorbate prior to illumination, and the metal can therefore no longer act as an electron acceptor; thus Q(A)(*)(-) is generated by illumination in such samples. (4) With dithionite, both the Cu(2+) and the quinone are reduced resulting in the presence of Q(A)(*)(-) in the dark. The suggested high redox potential of Cu(2+) when in the Fe(2+) site in PS II is in contrast to the situation in the bacterial reaction center where it has been shown in earlier work that the Cu(2+) is unreduced by dithionite. It cannot be ruled out however that Q(A)-Cu(2+) is formed and a magnetic interaction is responsible for the lack of the Q(A)(-) signal when no exogenous reductant is present. With this alternative possibility, the effects of reductants would be explained as the loss of Cu(2+) (due to formation of Cu(+)) leading to loss of the Cu(2+) from the Fe(2+) site due to the binding equilibrium. The quite different binding and redox behavior of the metal in the iron site in PS II compared to that of the bacterial reaction center is presumably a further reflection of the differences in the coordination of the iron in the two systems.  相似文献   

8.
Reactions of carbon monoxide with iron(II) diethyldithiocarbamate and iron(II) ethylxanthate were followed using solution IR spectroscopy. In DMF and CH3CN solutions, the only Fe—dithiocarbamate—carbon monoxide complex observed was cis-[Fe(CO)2(dedtc)2]. This complex formed rapidly and appeared to be very stable, resisting displacement of the coordinated CO molecules by other ligands. Fe(exa)2 showed very little coordination of CO in DMF solution, but in CH3CN solution formed the complex cis-[Fe(CO)2(exa)2] rapidly via the monocarbonyl intermediate [Fe(CO)(exa)2CH3CN]. In CHCl3 solution, in the presence of CO and added bases, a series of complexes, [Fe(CO)(exa)2L], where L = pyridine, pyrrolidine, diethylamine and triphenylphosphine, was formed. However, with the exception of [Fe(CO)(exa)23P)], these monocarbonyl complexes were unstable with respect to disproportionation to cis-[Fe(CO)2(exa)2] and [Fe(exa)2L2]. No mixed-ligand monocarbonyl complexes were observed with Fe(dedtc)2.  相似文献   

9.
Dorlet P  Rutherford AW  Un S 《Biochemistry》2000,39(26):7826-7834
The radical forms of two cofactors and an amino acid in the photosystem II (PS II) reaction center were studied by using high-field EPR both in frozen solution and in oriented multilayers. Their orientation with respect to the membrane was determined by using one-dimensionally oriented samples. The ring plane of the stable tyrosyl radical, Y(D)(*), makes an angle of 64 degrees +/- 5 degrees with the membrane plane, and the C-O direction is tilted by 72 degrees +/- 5 degrees in the plane of the radical compared to the membrane plane. The semiquinone, Q(A)(*)(-), generated by chemical reduction in samples lacking the non-heme iron, has its ring plane at an angle of 72 degrees +/- 5 degrees to the membrane plane, and the O-O axis is tilted by 21 degrees +/- 5 degrees in the plane of the quinone compared to the membrane plane. This orientation is similar to that of Q(A)(*)(-) in purple bacteria reaction centers except for the tilt angle which is slightly bigger. The pheophytin anion was generated by photoaccumulation under reducing conditions. Its ring plane is almost perpendicular to the membrane with an angle of 70 degrees +/- 5 degrees with respect to the membrane plane. This is very similar to the orientation of the pheophytin in purple bacteria reaction centers. The position of the g tensor with respect to the molecule is tentatively assigned for the anion radical on the basis of this comparison. In this work, the treatment of orientation data from EPR spectroscopy applied to one-dimensionally oriented multilayers is examined in detail, and improvements over previous approaches are given.  相似文献   

10.
CO(2) changes the biochemistry of peroxynitrite basically in two ways: (i) nitrating species is the CO(3)(-) / ()NO(2) radical pair, and (ii) peroxynitrite diffusion distance is significantly reduced. For peroxynitrite generated extracellularly this last effect is particularly dramatic at low cell density because CO(3)(-) and ()NO(2) are short-lived and decay mostly in the extracellular space or at the cell surface/membrane. This study was aimed to distinguish between peroxynitrite-induced extra- and intracellular modifications of red blood cells (RBC). Our results show that at low cell density and in the presence of CO(2) peroxynitrite induced the oxidation of surface thiols, the formation of 3-nitrotyrosine and DMPO-RBC adducts, and the down-regulation of glycophorins A and C (biomarkers of senescence). Reactivation of glycolysis reversed only the oxidation of surface thiols. Without CO(2) peroxynitrite also induced the oxidation of hemoglobin and glutathione, the accumulation of lactate, a decrease in ATP, the clustering of band 3, the externalization of phosphatidylserine, and the activation of caspases 8 and 3 (biomarkers of apoptosis). The latter biomarkers were all reversed by reactivation of glycolysis. We hypothesize that cell senescence could (generally) be derived by irreversible radical-mediated oxidation of membrane targets, while the appearance of apoptotic biomarkers could be bolstered by oxidation of intracellular targets. These results suggest that, depending on extracellular homolysis or diffusion to the intracellular space, peroxynitrite prompts RBCs toward either senescence or apoptosis through different oxidation mechanisms.  相似文献   

11.
Mo(CO)4(LL) complexes, where LL = polypyridyl ligands such as 2,2′-bipyridine and 1,10-phenanthroline, undergo quasi-reversible, one-electron oxidations in methylene chloride yielding the corresponding radical cations, [Mo(CO)4(LL)]+. These electrogenerated species undergo rapid ligand substitution in the presence of acetonitrile, yielding [Mo(CO)3(LL)(CH3CN)]+; rate constants for these substitutions were measured using chronocoulometry and were found to be influenced by the steric and electronic properties of the polypyridyl ligands. [Mo(CO)3(LL)(CH3CN)]+ radical cations, which could also be generated by reversible oxidation of Mo(CO)3(LL)(CH3CN) in acetonitrile, can be irreversibly oxidized yielding [Mo(CO)3(LL)(CH3CN)2]2+ after coordination by an additional acetonitrile. Infrared spectroelectrochemical experiments indicate the radical cations undergo ligand-induced net disproportionations that follow first-order kinetics in acetonitrile, ultimately yielding the corresponding Mo(CO)4(LL) and [Mo(CO)2(LL)(CH3CN)3]2+ species. Rate constants for the net disproportionation of [Mo(CO)3(LL)(CH3CN)]+ and the carbonyl substitution reaction of [Mo(CO)3(LL)(CH3CN)2]2+ were measured. Thin-layer bulk oxidation studies also provided infrared characterization data of [Mo(CO)4(ncp)]+ (ncp = neocuproine), [Mo(CO)3(LL)(CH3CN)]+, [Mo(CO)3(LL)(CH3CN)2]2+ and [Mo(CO)2(LL)(CH3CN)3]2+ complexes.  相似文献   

12.
Peroxynitrite (ONOO(-)/ONOOH), the product of the diffusion-limited reaction of nitric oxide (*NO) with superoxide (O(-*)(2)), has been implicated as an important mediator of tissue injury during conditions associated with enhanced *NO and O(-*)(2) production. Although several groups of investigators have demonstrated substantial oxidizing and cytotoxic activities of chemically synthesized peroxynitrite, others have proposed that the relative rates of *NO and production may be critical in determining the reactivity of peroxynitrite formed in situ (Miles, A. M., Bohle, D. S., Glassbrenner, P. A., Hansert, B., Wink, D. A., and Grisham, M. B. (1996) J. Biol. Chem. 271, 40-47). In the present study, we examined the mechanisms by which excess O(-*)(2) or *NO production inhibits peroxynitrite-mediated oxidation reactions. Peroxynitrite was generated in situ by the co-addition of a chemical source of *NO, spermineNONOate, and an enzymatic source of O(-*)(2), xanthine oxidase, with either hypoxanthine or lumazine as a substrate. We found that the oxidation of the model compound dihydrorhodamine by peroxynitrite occurred via the free radical intermediates OH and NO(2), formed during the spontaneous decomposition of peroxynitrite and not via direct reaction with peroxynitrite. The inhibitory effect of excess O(-*)(2) on the oxidation of dihydrorhodamine could not be ascribed to the accumulation of the peroxynitrite scavenger urate produced from the oxidation of hypoxanthine by xanthine oxidase. A biphasic oxidation profile was also observed upon oxidation of NADH by the simultaneous generation of *NO and O(-*)(2). Conversely, the oxidation of glutathione, which occurs via direct reaction with peroxynitrite, was not affected by excess production of *NO. We conclude that the oxidative processes initiated by the free radical intermediates formed from the decomposition of peroxynitrite are inhibited by excess production of *NO or O(-*)(2), whereas oxidative pathways involving a direct reaction with peroxynitrite are not altered. The physiological implications of these findings are discussed.  相似文献   

13.
Single channel and macroscopic current recording was used to investigate block of the cystic fibrosis transmembrane conductance regulator (CFTR) Cl(-) channel pore by the permeant anion Au(CN)2(-). Block was 1-2 orders of magnitude stronger when Au(CN)2(-) was added to the intracellular versus the extracellular solution, depending on membrane potential. A point mutation within the pore, T-338A, strongly decreased the asymmetry of block, by weakening block by intracellular Au(CN)2(-) and at the same time strengthening block by external Au(CN)2(-). Block of T-338A, but not wild-type, was strongest at the current reversal potential and weakened by either depolarization or hyperpolarization. In contrast to these effects, the T-338A mutation had no impact on block by the impermeant Pt(NO2)4(2-) ion. We suggest that the CFTR pore has at least two anion binding sites at which Au(CN)2(-) and Pt(NO2)4(2-) block Cl- permeation. The T-338A mutation decreases a barrier for Au(CN)2(-) movement between different sites, leading to significant changes in its blocking action. Our finding that apparent blocker binding affinity can be altered by mutagenesis of a residue which does not contribute to a blocker binding site has important implications for interpreting the effects of mutagenesis on channel blocker effects.  相似文献   

14.
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   

15.
The carbonate radical anion (CO(3)) is believed to be an important intermediate oxidant derived from the oxidation of bicarbonate anions and nitrosoperoxocarboxylate anions (formed in the reaction of CO(2) with ONOO(-)) in cellular environments. Employing nanosecond laser flash photolysis methods, we show that the CO(3) anion can selectively oxidize guanines in the self-complementary oligonucleotide duplex d(AACGCGAATTCGCGTT) dissolved in air-equilibrated aqueous buffer solution (pH 7.5). In these time-resolved transient absorbance experiments, the CO(3) radicals are generated by one-electron oxidation of the bicarbonate anions (HCO(3)(-)) with sulfate radical anions (SO(4)) that, in turn, are derived from the photodissociation of persulfate anions (S(2)O(8)(2-)) initiated by 308-nm XeCl excimer laser pulse excitation. The kinetics of the CO(3) anion and neutral guanine radicals, G(-H)( small middle dot), arising from the rapid deprotonation of the guanine radical cation, are monitored via their transient absorption spectra (characteristic maxima at 600 and 315 nm, respectively) on time scales of microseconds to seconds. The bimolecular rate constant of oxidation of guanine in this oligonucleotide duplex by CO(3) is (1.9 +/- 0.2) x 10(7) m(-1) s(-1). The decay of the CO(3) anions and the formation of G(-H)( small middle dot) radicals are correlated with one another on the millisecond time scale, whereas the neutral guanine radicals decay on time scales of seconds. Alkali-labile guanine lesions are produced and are revealed by treatment of the irradiated oligonucleotides in hot piperidine solution. The DNA fragments thus formed are identified by a standard polyacrylamide gel electrophoresis assay, showing that strand cleavage occurs at the guanine sites only. The biological implications of these oxidative processes are discussed.  相似文献   

16.
To elucidate potential mechanisms of S-nitrosothiol formation in vivo, we studied nitrosation of GSH and albumin by nitric oxide ((*)NO), peroxynitrite, and (*)NO/O(2)(*)(-). In the presence of O(2), (*)NO yielded 20% of S-nitrosoglutathione (GSNO) at pH 7.5. Ascorbate and the spin trap 4-hydroxy-[2,2,4,4-tetramethyl-piperidine-1-oxyl] (TEMPOL) inhibited GSNO formation by 67%. Electron paramagnetic resonance spectroscopy with 5-diethoxyphosphoryl-5-methyl-1-pyrroline-N-oxide (DEPMPO) demonstrated intermediate formation of glutathionyl radicals, suggesting that GSNO formation by (*)NO/O(2) is predominantly mediated by (*)NO(2). Peroxynitrite-triggered GSNO formation (0.06% yield) was stimulated 10- and 2-fold by ascorbate and TEMPOL, respectively. Co-generation of (*)NO and O(2)(*)(-) at equal fluxes yielded less GSNO than (*)NO alone, but was 100-fold more efficient (8% yield) than peroxynitrite. Moreover, in contrast to the reaction of peroxynitrite, GSNO formation by (*)NO/O(2)(*)(-) was inhibited by ascorbate. Similar results were obtained with albumin instead of GSH. We propose that sulfhydryl compounds react with O(2)(*)(-) to initiate a chain reaction that forms radical intermediates which combine with (*)NO to yield GSNO. In RAW 264.7 macrophages, S-nitrosothiol formation by (*)NO/O(2) and (*)NO/O(2)(*)(-) occurred with relative efficiencies comparable to those in solution. Our results indicate that concerted generation of (*)NO and O(2)(*)(-) may essentially contribute to nitrosative stress in inflammatory diseases.  相似文献   

17.
Enhanced absorption is observed in the (15)N NMR spectra of (15)NO(-)(3) during decomposition of peroxynitrite and the peroxynitrite-CO(2) adduct at pH 5.25, indicating the formation of (15)NO(-)(3) in radical pairs [(15)NO(*)(2), HO(*)] and [(15)NO(*)(2), CO(*-)(3)]. During the reaction of peroxynitrite and the peroxynitrite-CO(2) adduct with L-tyrosine, the (15)N NMR signal of the nitration product 3-nitrotyrosine exhibits emission showing a radical pathway of its formation. The nuclear polarization is built up in radical pairs [(15)NO(*)(2), tyr(*)] generated by free radical encounters of nitrogen dioxide and tyrosinyl radicals. The (15)N NMR signal of (15)NO(-)(2) formed during reaction of peroxynitrite with L-tyrosine appears in emission. It is concluded that tyrosinyl radicals are generated by reaction of nitrogen dioxide with L-tyrosine. In contrast to this, (15)NO(-)(2) does not show (15)N chemically induced dynamic nuclear polarization (CIDNP) during reaction of the peroxynitrite-CO(2) adduct with L-tyrosine, indicating a different reaction mechanism, which is assumed to be a hydrogen transfer between CO(*-)(3) and L-tyrosine. Emission is also observed in the (15)N NMR signals of 2-nitro-4-fluorophenol, 3-nitro-4-hydroxyphenylacetic acid, 2-nitrophenol, and 4-nitrophenol during reaction of 4-fluorophenol, 4-hydroxyphenylacetic acid, and phenol with peroxynitrite and the peroxynitrite-CO(2) adduct. 3-Nitro-4-hydroxyphenylacetic acid is also observed in emission during reaction of phenylacetic acid with peroxynitrite, but is not formed with the peroxynitrite-CO(2) adduct. The magnitude of the (15)N CIDNP effect during reaction of peroxynitrite with 4-fluorophenol and of the peroxynitrite-CO(2) adduct with 4-fluorophenol and phenol is determined. It excludes the occurrence of nonradical reactions. Only weak emission signals are observed during the reaction of peroxynitrite with phenol in (15)NO(-)(2), 2-nitrophenol, and 4-nitrophenol. 2-Nitrophenol is only formed in traces, and 4-nitrophenol is only formed in higher yields. The latter might be generated in part via a nonradical pathway.  相似文献   

18.
Chain scission of hyaluronan by peroxynitrite   总被引:2,自引:0,他引:2  
The reaction of peroxynitrite with the biopolymer hyaluronan has been studied using stopped-flow techniques combined with detection of molecular weight changes using the combination of gel permeation chromatography and multiangle laser light scattering. From the effect of peroxynitrite on the yield of hyaluronan chain breaks, it was concluded that the chain breaks were caused by hydroxyl radicals which escape a cage containing the *OH NO*(2) radical pair. The yield of free hydroxyl radicals was determined as 5+/-1% (as a proportion of the total peroxynitrite concentration). At high peroxynitrite concentrations, it was observed that the yield of chain breaks leveled out, an effect largely attributable to the scavenging of hydroxyl radicals by nitrite ions present in the peroxynitrite preparation. These experiments also provided some support for a previous proposal that the adduct formed between ONOOH and ONOO(-) might itself produce hydroxyl radicals. The rate of this reaction would have to be of the order of 0.05 s(-1) to produce hydroxyl radical yields that would account quantitatively for chain break yields at high peroxynitrite concentrations. By carrying out experiments at higher hyaluronan concentrations, it was also concluded that an additional yield of chain breaks was produced by the bimolecular reaction of the polymer with ONOOH at a rate constant of about 10 dm(3)mol(-1)s(-1). At 5.3 x 10(-3)mol dm(-3) hyaluronan, this amounted to 3.5% chain breaks (per peroxynitrite concentration). These conclusions support the proposal that the yield of hydroxyl radicals arising from the isomerization of ONOOH to nitrate ions is relatively low.  相似文献   

19.
A comparative study of the rates of ferrocyanide-catalyzed oxidation of several oxymyoglobins by molecular oxygen is reported. Oxidation of the native oxymyoglobins from sperm whale, horse and pig, as well as the chemically modified (MbO(2)) sperm whale oxymyoglobin, with all accessible His residues alkylated by sodium bromoacetate (CM-MbO(2)), and the mutant sperm whale oxymyoglobin [MbO(2)(His119-->Asp)], was studied. The effect of pH, ionic strength and the concentration of anionic catalyst ferrocyanide, [Fe(CN)(6)](4-), on the oxidation rate is investigated, as well as the effect of MbO(2) complexing with redox-inactive Zn(2+), which forms the stable chelate complex with functional groups of His119, Lys16 and Asp122, all located nearby. The catalytic mechanism was demonstrated to involve specific [Fe(CN)(6)](4-) binding to the protein in the His119 region, which agrees with a high local positive electrostatic potential and the presence of a cavity large enough to accommodate [Fe(CN)(6)](4-) in that region. The protonation of the nearby His113 and especially His116 plays a very important role in the catalysis, accelerating the oxidation rate of bound [Fe(CN)(6)](4-) by dissolved oxygen. The simultaneous occurrence of both these factors (i.e. specific binding of [Fe(CN)(6)](4-) to the protein and its fast reoxidation by oxygen) is necessary for the efficient ferrocyanide-catalyzed oxidation of oxymyoglobin.  相似文献   

20.
Heme oxygenase (HO) catalyzes heme degradation by utilizing O(2) and reducing equivalents to produce biliverdin IX alpha, iron, and CO. To avoid product inhibition, the heme[bond]HO complex (heme[bond]HO) is structured to markedly increase its affinity for O(2) while suppressing its affinity for CO. We determined the crystal structures of rat ferrous heme[bond]HO and heme[bond]HO bound to CO, CN(-), and NO at 2.3, 1.8, 2.0, and 1.7 A resolution, respectively. The heme pocket of ferrous heme-HO has the same conformation as that of the previously determined ferric form, but no ligand is visible on the distal side of the ferrous heme. Fe[bond]CO and Fe[bond]CN(-) are tilted, whereas the Fe[bond]NO is bent. The structure of heme[bond]HO bound to NO is identical to that bound to N(3)(-), which is also bent as in the case of O(2). Notably, in the CO- and CN(-)-bound forms, the heme and its ligands shift toward the alpha-meso carbon, and the distal F-helix shifts in the opposite direction. These shifts allow CO or CN(-) to bind in a tilted fashion without a collision between the distal ligand and Gly139 O and cause disruption of one salt bridge between the heme and basic residue. The structural identity of the ferrous and ferric states of heme[bond]HO indicates that these shifts are not produced on reduction of heme iron. Neither such conformational changes nor a heme shift occurs on NO or N(3)(-) binding. Heme[bond]HO therefore recognizes CO and O(2) by their binding geometries. The marked reduction in the ratio of affinities of CO to O(2) for heme[bond]HO achieved by an increase in O(2) affinity [Migita, C. T., Matera, K. M., Ikeda-Saito, M., Olson, J. S., Fujii, H., Yoshimura, T., Zhou, H., and Yoshida, T. (1998) J. Biol. Chem. 273, 945-949] is explained by hydrogen bonding and polar interactions that are favorable for O(2) binding, as well as by characteristic structural changes in the CO-bound form.  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