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 750 毫秒
1.
The peroxidase-like activity of cytochrome c is considerably increased by unfolding of the protein. The enhancement of the activity is due to the higher reaction rate of unfolded cytochrome c with hydrogen peroxide, which is the rate-determining step in the peroxidase cycle of cytochrome c (Gebicka, L., 2001, Res Chem Intermed 27, 717-23). In this study we checked whether combined action of two unfolding factors, SDS and peroxynitrite or radiation (hydroxyl radicals), increases the peroxidase-like activity of cytochrome c more than any single treatment alone. Peroxynitrite reacts with SDS-modified cytochrome c in the same way as with native cytochrome c, via intermediate radical products, *OH/*NO2, arising from peroxynitrite homolysis. We found that SDS-modified cytochrome c is much more sensitive to oxidative damage than the native protein. Partial unfolding of cytochrome c by SDS causes the peroxide substrate to have a better access to the heme center. On the other hand, the amino acids located in the vicinity of the active site and/or heme group become accessible for oxidizing radicals. The overall effect observed is that the peroxidase-like activity of SDS-modified cytochrome c decreases with an increase of the concentration of the oxidizing species (peroxynitrite or radiolytically generated hydroxyl radicals). The damage of SDS-modified cytochrome c caused by irradiation is much more significant than that observed after peroxynitrite treatment.  相似文献   

2.
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   

3.
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   

4.
NAD(P)H acts as a two-electron reductant in physiological, enzyme-controlled processes. Under nonenzymatic conditions, a couple of one-electron oxidants easily oxidize NADH to the NAD(.) radical. This radical reduces molecular oxygen to the superoxide radical (O-(2)) at a near to the diffusion-controlled rate, thereby subsequently forming hydrogen peroxide (H(2)O(2)). Because peroxynitrite can act as a one-electron oxidant, the reaction of NAD(P)H with both authentic peroxynitrite and the nitric oxide ((. )NO) and O-(2) releasing compound 3-morpholinosydnonimine N-ethylcarbamide (SIN-1) was studied. Authentic peroxynitrite oxidized NADH with an efficiency of approximately 25 and 8% in the absence and presence of bicarbonate/carbon dioxide (HCO(3)(-)/CO(2)), respectively. NADH reacted 5-100 times faster with peroxynitrite than do the known peroxynitrite scavengers glutathione, cysteine, and tryptophan. Furthermore, NADH was found to be highly effective in suppressing peroxynitrite-mediated nitration reactions even in the presence of HCO(3)(-)/CO(2). Reaction of NADH with authentic peroxynitrite resulted in the formation of NAD(+) and O-(2) and, thus, of H(2)O(2) with yields of about 3 and 10% relative to the added amounts of peroxynitrite and NADH, respectively. Peroxynitrite generated in situ from SIN-1 gave virtually the same results; however, two remarkable exceptions were recognized. First, the efficiency of NADH oxidation increased to 60-90% regardless of the presence of HCO(3)(-)/CO(2), along with an increase of H(2)O(2) formation to about 23 and 35% relative to the amounts of added SIN-1 and NADH. Second, and more interesting, the peroxynitrite scavenger glutathione (GSH) was needed in a 75-fold surplus to inhibit the SIN-1-dependent oxidation of NADH half-maximal in the presence of HCO(3)(-)/CO(2). Similar results were obtained with NADPH. Hence, peroxynitrite or radicals derived from it (such as, e.g. the bicarbonate radical or nitrogen dioxide) indeed oxidize NADH, leading to the formation of NAD(+) and, via O-(2), of H(2)O(2). When peroxynitrite is generated in situ in the presence of HCO(3)(-)/CO(2), i.e. under conditions mimicking the in vivo situation, NAD(P)H effectively competes with other known scavengers of peroxynitrite.  相似文献   

5.
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   

6.
Hemoproteins are one of the major targets of peroxynitrite in vivo. It has been proposed that the bimolecular heme/peroxynitrite interaction results in both peroxynitrite inactivation (scavenging) and catalysis of tyrosine nitration. In this study, we used spectroscopic techniques to analyze the reaction of peroxynitrite with human methemoglobin (metHb). Although conventional differential spectroscopy did not reveal heme changes, our results suggest that, in the absence of bicarbonate, the heme in metHb reacts bimolecularly with peroxynitrite but is quickly back-reduced by the reaction products. This hypothesis is based on two indirect observations. First, metHb prevents the peroxynitrite-mediated nitration of a target dipeptide, Ala-Tyr, and second, it promotes the isomerization of peroxynitrite to nitrate. Both the scavenging and the isomerization activities of metHb were heme-dependent and inhibited by CO(2). Ferrous cytochrome c was an efficient scavenger of peroxynitrite, but in the ferric form did not show either scavenging or isomerization activities. We found no evidence of an increase in Ala-Tyr nitration with these hemoproteins. Peroxynitrite-treated metHb induced the formation of a long-lived radical assigned to tyrosine by spin-trapping studies. This radical, however, did not allow us to predict an interaction of peroxynitrite with heme. Hb was nitrated by peroxynitrite/CO(2) mainly in tyrosines beta 130, alpha 42, and alpha 140 and, to a lesser extent, alpha 24. The nitration of alpha chain tyrosines more exposed to the solvent (alpha 140 and alpha 24) was higher in CO-Hb and metHb, while nitration of alpha 42, the tyrosine nearest to the heme, was higher in oxyHb. We deduce that the heme/peroxynitrite interaction, which is inhibited in CO-Hb and metHb, affects alpha tyrosine nitration in two opposite ways, i.e., by protecting exposed residues and by promoting nitration of the residue nearest to the heme. Conversely, nitration of beta Tyr 130 was comparable in oxyHb, metHb, and CO-Hb, suggesting a mechanism involving only nitrating species formed during peroxynitrite decay.  相似文献   

7.
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   

8.
In this study, we irradiated the antioxidant kaempferol in ethanol and methanol solutions with gamma rays at doses ranging from 0.2-20 kGy. NMR and ES-MS spectroscopy were used to identify radiolysis products. Two depsides, [2-[(4'-hydroxybenzoyl)oxy]-4,6-dihydroxyphenyl](oxo) methyl acetate and [2-[(4'-hydroxybenzoyl)oxy]-4,6-dihydroxyphenyl](oxo) ethyl acetate, were the major compounds of kaempferol degradation in methanol and in ethanol, respectively. Other products formed in low concentrations were identified as [4-hydroxyphenyl](oxo) methyl acetate, [4-hydroxyphenyl](oxo) ethyl acetate, and depside [2-[(4'-hydroxybenzoyl)oxy]-4,6-dihydroxyphenyl](oxo) acetic acid. The formation of the latter was observed in both solvents. We propose degradation mechanisms that suggest that (.)CH(2)OH and CH(3)(.)CHOH, produced by solvent radiolysis, react with the 3-OH kaempferol group because of its high H-donor capacity. pi-Electron delocalization in the flavonoxy formed after the first H-transfer leads to C-ring opening and consequently to the formation of depsides. G calculation of the degradation products and of (.)CH(2)OH and CH(3)(.)CHOH radicals confirmed the proposed mechanism of kaempferol radiolysis. The rate constants for the reaction between kaempferol and these free radicals were also calculated. Formation of depside has also been observed in many studies of the oxidation of flavonoids; those studying human metabolism have suggested similar redox transformation of flavonols. The antioxidant activities of radiolysis products were evaluated and compared to those of kaempferol.  相似文献   

9.
1. Techniques and experiments are described concerned with the millisecond kinetics of EPT-detectable changes brought about in cytochrome c oxidase by reduced cytochrome c and, after reduction with various agents, by reoxidation with O2 or ferricyanide. Some experiments in the presence of ligands are also reported. Light absorption was monitored by low-temperature reflectance spectroscopy. 2. In the rapid phase of reduction of cytochrome c oxidase by cytochrome c (less than 50 ms) approx. 0.5 electron equivalent per heme a is transferred mainly to the low-spin heme component of cytochrome c oxidase and partly to the EPR-detectable copper. In a slow phase (less than 1 s) the copper is reoxidized and high-spin ferric heme signals appear with a predominant rhombic component. Simultaneously the absorption band at 655 nm decreases and the Soret band at 444 nm appears between the split Soret band (442 and 447 nm) of reduced cytochrome a. 3. On reoxidation of reduced enzyme by oxygen all EPR and optical features are restored within 6 ms. On reoxidation by O2 in the presence of an excess of reduced cytochrome c, states can be observed where the low-spin heme and copper signals are largely absent but the absorption at 655 nm is maximal, indicating that the low-spin heme and copper components are at the substrate side and the component(s) represented in the 655 nm absorption at the O2 side of the system. On reoxidation with ferricyanide the 655 nm absorption is not readily restored but a ferric high-spin heme, represented by a strong rhombic signal, accumulates. 4. On reoxidation of partly reduced enzyme by oxygen, the rhombic high-spin signals disappear within 6 ms., whereas the axial signals disappear more slowly, indicating that these species are not in rapid equilibrium. Similar observations are made when partly reduced enzyme is mixed with CO. 5. The results of this and the accompanying paper are discussed and on this basis an assignment of the major EPR signals and of the 655 nm absorption is proposed, which in essence is that published previously (Hartzell, C.R., Hansen, R.E. and Beinert, H. (1973) Proc. Natl. Acad. Sci. U.S. 70, 2477-2481). Both the low-spin (g=o; 2.2; 1.5) and slowly appearing high-spin (g=6; 2) signals are attributed to ferric cytochrome a, whereas the 655 nm absorption is thought to arise from ferric cytochrome a3, when it is present in a state of interaction with EPR-undectectable copper. Alternative possibilities and possible inconsistencies with this proposal are discussed.  相似文献   

10.
In a series of heme and non-heme proteins the nitration of tyrosine residues was assessed by complete pronase digestion and subsequent HPLC-based separation of 3-nitrotyrosine. Bolus addition of peroxynitrite caused comparable nitration levels in all tested proteins. Nitration mainly depended on the total amount of tyrosine residues as well as on surface exposition. In contrast, when superoxide and nitrogen monoxide (NO) were generated at equal rates to yield low steady-state concentrations of peroxynitrite, metal catalysis seemed to play a dominant role in determining the sensitivity and selectivity of peroxynitrite-mediated tyrosine nitration in proteins. Especially, the heme-thiolate containing proteins cytochrome P450(BM-3) (wild type and F87Y variant) and prostacyclin synthase were nitrated with high efficacy. Nitration by co-generated NO/O(2)(-) was inhibited in the presence of superoxide dismutase. The NO source alone only yielded background nitration levels. Upon changing the NO/O(2)(-) ratio to an excess of NO, a decrease in nitration in agreement with trapping of peroxynitrite and derived radicals by NO was observed. These results clearly identify peroxynitrite as the nitrating species even at low steady-state concentrations and demonstrate that metal catalysis plays an important role in nitration of protein-bound tyrosine.  相似文献   

11.
Reduction of cytochrome P-450S21 (SF) (SF, substrate-free; purified from bovine adrenocortical microsomes) with sodium dithionite (Na2S2O4) in the presence of phenylisocyanide produced a ferrous cytochrome P-450S21 (SF)-phenylisocyanide complex with Soret absorbance maxima at 429 and 456 nm. On the other hand, when a preformed ferric cytochrome P-450S21 (SF)-NADPH-cytochrome-P-450 reductase (Fp2) complex was reduced chemically or enzymatically under the same conditions, the absorbance spectrum of the ferrous cytochrome P-450S21 (SF)-phenylisocyanide complex changed drastically, as characterized by an increase in absorbance intensity at 429 nm and a decrease at 456 nm. Similar spectral changes were observed by addition of reduced Fp2 to the preformed ferrous cytochrome P-450S21 (SF)-phenylisocyanide complex. Experiments to reduce a ferric cytochrome P-450S21 (SF)-phenylisocyanide complex with sodium dithionite in the presence of various amounts of Fp2 showed that; (1), the spectral change reached maxima for both absorption increase at 429 nm and decrease at 456 nm when cytochrome P-450S21 and Fp2 were previously mixed at the cytochrome P-450S21:Fp2 ratio of 1:5; (2), the spectral change was suppressed in 300 mM potassium phosphate buffer (pH 7.4). These results suggest that the absorbance spectral change is due to a conformational change around the heme moiety induced by association with reduced Fp2.  相似文献   

12.
Escherichia coli were transformed by electroporation to introduce a plasmid harboring a GFP gene-containing vector. The fluorescence of the purified GFP isolated from the transformant was quenched by myeloperoxidase (MPO)-generated HOCl, by peroxynitrous acid (ONOOH) and by enzymatically or radiolytically generated NO(2)(.) but not by other putative neutrophil-generated oxidants. Fluorescence from the bacterium was effectively quenched by HOCl but not peroxynitrite, oxidizing radicals derived from its O-O bond homolysis, or the other oxidants under study. Exposure of serum-opsonized bacteria to human neutrophils resulted in extensive loss of GFP fluorescence; fluorescence microscopy revealed that phagocytosed bacteria were completely quenched but that bacteria remaining in the external media were unquenched. Addition of sodium azide to the medium to inhibit MPO prevented neutrophil-mediated fluorescence quenching. Because the amount of HOCl required to inhibit bacterial fluorescence was an order of magnitude greater than required to inhibit colonial growth, these results imply that sufficient HOCl was formed within the neutrophil phagosome to kill the microbe.  相似文献   

13.
Heme oxygenase-2 (HO-2) is the constitutive cognate of the heat-shock protein-32 family of proteins. These proteins catalyze oxidative cleavage of heme to CO and biliverdin, and release Fe. HO-2 is a hemoprotein and binds heme at heme regulatory motifs (HRMs) with a conserved Cys-Pro pair; two copies of HRM are present in HO-2 (Cys264 and Cys281). The HO-2 HRMs are not present in HO-1 and are not involved in HO-2 catalytic activity. Optical CD, and spectral and activity analyses were used to examine reactivity of HO isozymes with NO species produced by NO donors. Purified Escherichia coli-expressed HO preparations, wild-type HO-2, Cys264/Cys281 --> Ala/Ala HO-2-mutant (HO-2-mut) and HO-1 preparations were used. A type II change (red shift) of the Soret band (405 nm --> 413-419 nm) was observed when wild-type HO-2 was treated with sodium nitroprusside (SNP), S-nitroglutathione (GSNO), S-nitroso-N-acetylpenicillamine (SNAP) or 3-morpholinosydnonimine (SIN-1); the NO scavenger, hydroxocobalamin (HCB) prevented the shift. Only SIN-1, which produces peroxynitrite by generating both NO and superoxide anion, decreased the Soret region absorption and the pyridine hemochromogen spectrum of HO-2; superoxide dismutase (SOD) blocked the decrease. Binding of heme to HO-2 protein was required for shift and/or decrease in absorption of the Soret band. NO donors significantly inhibited HO-2 activity, with SNP being the most potent inhibitor (> 40%). Again, trapping NO with HCB blocked HO-2 inactivation. HO-1 and HO-2-mut were not inactivated by NO donors. CD data suggest that the decrease in HO-2 activity was not related to change by NO species of the secondary structure of HO-2. Western blot analysis suggests that NO donors did not cause HO-1 protein loss and Northern blot analysis of HeLa cells treated with SIN-1 and SNP indicates that, unlike HO-1 mRNA, which is remarkably responsive to the treatments, HO-2 mRNA levels were modestly increased ( approximately two to threefold) by NO donors. The data are consistent with the possibility that NO interaction with HO-2-bound heme effects electronic interactions of residues involved in substrate binding and/or oxygen activation. The findings permit the hypothesis that HO-2 and NO are trans-inhibitors, whereby biological activity of NO is attenuated by interaction with HO-2, serving as an intracellular 'sink' for the heme ligand, and NO inhibits HO-2 catalytic activity. As such, the cellular level of both signaling molecules, CO and NO would be moderated.  相似文献   

14.
This work compares the effect of photogenerated singlet oxygen (O(2)((1)Delta(g))) (type II mechanism) and free radicals (type I mechanism) on cytochrome c structure and reactivity. Both reactive species were obtained by photoexcitation of methylene blue (MB(+)) in the monomer and dimer forms, respectively. The monomer form is predominant at low dye concentrations (up to 8 microm) or in the presence of an excess of SDS micelles, while dimers are predominant at 0.7 mm SDS. Over a pH range in which cytochrome c is in the native form, O(2) ((1)Delta(g)) and free radicals induced a Soret band blue shift (from 409 to 405 nm), predominantly. EPR measurements revealed that the blue shift of the Soret band was compatible with conversion of the heme iron from its native low spin state to a high spin state with axial symmetry (g approximately 6.0). Soret band bleaching, due to direct attack on the heme group, was only detected under conditions that favored free radical production (MB(+) dimer in SDS micelles) or in the presence of a less structured form of the protein (above pH 9.3). Matrix-assisted laser desorption ionization time-of-flight mass spectrometry of the heme group and the polypeptide chain of cytochrome c with Soret band at 405 nm (cytc405) revealed no alterations in the mass of the cytc405 heme group but oxidative modifications on methionine (Met(65) and Met(80)) and tyrosine (Tyr(74)) residues. Damage of cytc405 tyrosine residue impaired its reduction by diphenylacetaldehyde, but not by beta-mercaptoethanol, which was able to reduce cytc405, generating cytochrome c Fe(II) in the high spin state (spin 2).  相似文献   

15.
Soluble guanylate cyclase (sGC) is a heterodimeric hemoprotein that catalyzes the conversion of GTP to cGMP. Upon binding NO to its heme cofactor, purified sGC was activated 300-fold. sGC was only activated 67-fold by nitroglycerin (GTN) and Cys; and in the absence of Cys, GTN did not activate sGC. Electronic absorption spectroscopy studies showed that upon NO binding, the Soret of ferrous sGC shifted from 431 to 399 nm. The data also revealed that activation of sGC by GTN/Cys was not via the expected ferrous heme-NO species as indicated by the absence of the 399 nm heme Soret. Furthermore, EPR studies of the reaction of GTN/Cys with sGC confirmed that no ferrous heme-NO species was formed but that there was heme oxidation. Potassium ferricyanide is known to oxidize ferrous sGC to the ferric oxidation state. Spectroscopic and activity data for the reactions of sGC with GTN alone or with K(3)Fe(CN)(6) were indistinguishable. These data suggest the following: 1) GTN/Cys do not activate sGC via GTN biotransformation to NO in vitro, and 2) in the absence of added thiol, GTN oxidizes sGC.  相似文献   

16.
Huang L  Abu-Soud HM  Hille R  Stuehr DJ 《Biochemistry》1999,38(6):1912-1920
The neuronal NO synthase (nNOS) heme binds self-generated NO, and this negatively regulates NO synthesis. Here we utilized the nNOS oxygenase domain and full-length nNOS along with various spectroscopic methods to (1) study formation of the six-coordinate ferrous NO complex and its conversion to a five-coordinate NO complex and (2) investigate the spectral and catalytic properties of the five-coordinate NO complex following its air oxidation to a ferric enzyme. NO bound quickly to ferrous nNOS oxygenase to form a six-coordinate NO complex (kon and koff values of 1.25 x 10(-)3 mM-1 s-1 and 128 s-1 at 10 degreesC, respectively) that was stable in the presence of L-arginine or tetrahydrobiopterin (BH4) but was converted to a five-coordinate NO complex in a biphasic process (k = 0.1 and 0.01 s-1 at 10 degreesC) in the absence of these molecules. Air oxidation of the ferrous six-coordinate NO complex generated an enzyme with full activity and ferrous-CO Soret absorbance at 444 nm. In contrast, oxidation of the five-coordinate NO complex generated an inactive dimer with ferrous-CO Soret absorbance at 420 nm, indicating nNOS was converted to a ferric P420 form. Incubation of ferric P420 nNOS with BH4 alone or BH4 and L-arginine resulted in time-dependent reactivation of catalysis and associated recovery of P450 character. Thus, nNOS is a heme-thiolate protein that can undergo a reversible P450-P420 conversion. BH4 has important roles in preventing P420 formation during NO synthesis, and in rescuing P420 nNOS.  相似文献   

17.
The UV-visible absorption and magnetic circular dichroism (MCD) spectra of the ferric, ferrous, CO-ligated forms and kinetic photolysis intermediates of the tetraheme electron-transfer protein cytochrome c3 (Cc3) are reported. Consistent with bis-histidinyl axial coordination of the hemes in this Class III c-type cytochrome, the Soret and visible region MCD spectra of ferric and ferrous Cc3 are very similar to those of other bis-histidine axially coordinated hemeproteins such as cytochrome b5. The MCD spectra indicate low spin state for both the ferric (S = 1/2) and ferrous (S = 0) oxidation states. CO replaces histidine as the axial sixth ligand at each heme site, forming a low-spin complex with an MCD spectrum similar to that of myoglobin-CO. Photodissociation of Cc3-CO (observed photolysis yield = 30%) produces a transient five-coordinate, high-spin (S = 2) species with an MCD spectrum similar to deoxymyoglobin. The recombination kinetics of CO with heme Fe are complex and appear to involve at least five first-order or pseudo first-order rate processes, corresponding to time constants of 5.7 microseconds, 62 microseconds, 425 microseconds, 2.9 ms, and a time constant greater than 1 s. The observed rate constants were insensitive to variation of the actinic photon flux, suggesting noncooperative heme-CO rebinding. The growing in of an MCD signal characteristic of bis-histidine axial ligation within tens of microseconds after photodissociation shows that, although heme-CO binding is thermodynamically favored at 1 atm CO, binding of histidine to the sixth axial site competes kinetically with CO rebinding.  相似文献   

18.
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   

19.
The mechanism of the reactions of myoglobin and hemoglobin with *OH and CO3*- in the presence of oxygen was studied using pulse and gamma-radiolysis. Unlike *NO2, which adds to the porphyrin iron, *OH and CO3*- form globin radicals. These secondary radicals oxidize the Fe(II) center through both intra- and intermolecular processes. The intermolecular pathway was further demonstrated when BSA radicals derived from *OH or CO3*- oxidized oxyhemoglobin and oxymyoglobin to their respective ferric states. The oxidation yields obtained by pulse radiolysis were lower compared to gamma-radiolysis, where the contribution of radical-radical reactions is negligible. Full oxidation yields by *OH-derived globin radicals could be achieved only at relatively high concentrations of the heme protein mainly via an intermolecular pathway. It is suggested that CO3*- reaction with the protein yields Tyr and/or Trp-derived phenoxyl radicals, which solely oxidize the porphyrin iron under gamma-radiolysis conditions. The *OH particularly adds to aromatic residues, which can undergo elimination of H2O forming the phenoxyl radical, and/or react rapidly with O2 yielding peroxyl radicals. The peroxyl radical can oxidize a neighboring porphyrin iron and/or give rise to superoxide, which neither oxidize nor reduce the porphyrin iron. The potential physiological implications of this chemistry are that hemoglobin and myoglobin, being present at relatively high concentrations, can detoxify highly oxidizing radicals yielding the respective ferric states, which are not toxic.  相似文献   

20.
Liszkay A  van der Zalm E  Schopfer P 《Plant physiology》2004,136(2):3114-23; discussion 3001
Cell extension in the growing zone of plant roots typically takes place with a maximum local growth rate of 50% length increase per hour. The biochemical mechanism of this dramatic growth process is still poorly understood. Here we test the hypothesis that the wall-loosening reaction controlling root elongation is effected by the production of reactive oxygen intermediates, initiated by a NAD(P)H oxidase-catalyzed formation of superoxide radicals (O(2)(.-)) at the plasma membrane and culminating in the generation of polysaccharide-cleaving hydroxyl radicals ((.)OH) by cell wall peroxidase. The following results were obtained using primary roots of maize (Zea mays) seedlings as experimental material. (1) Production of O(2)(.-), H(2)O(2), and (.)OH can be demonstrated in the growing zone using specific histochemical assays and electron paramagnetic resonance spectroscopy. (2) Auxin-induced inhibition of growth is accompanied by a reduction of O(2)(.-) production. (3) Experimental generation of (.)OH in the cell walls with the Fenton reaction causes wall loosening (cell wall creep), specifically in the growing zone. Alternatively, wall loosening can be induced by (.)OH produced by endogenous cell wall peroxidase in the presence of NADH and H(2)O(2). (4) Inhibition of endogenous (.)OH formation by O(2)(.-) or (.)OH scavengers, or inhibitors of NAD(P)H oxidase or peroxidase activity, suppress elongation growth. These results show that juvenile root cells transiently express the ability to generate (.)OH, and to respond to (.)OH by wall loosening, in passing through the growing zone. Moreover, inhibitor studies indicate that (.)OH formation is essential for normal root growth.  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