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 78 毫秒
1.
Enzymatic oxidation of dipyridamole (DIP) by horseradish peroxidase-hydrogen peroxide system (HRP-H2O2) in aqueous and micellar solutions was carried out. The reaction was monitored by optical absorption and fluorescence techniques. In aqueous solution at pH 7.0 and pH 9.0, the disappearance of the characteristic bands of DIP centered at 400 nm and 280 nm was observed. A new strong band at 260 nm is observed for the oxidation product(s) with shoulders at 322 nm and 390 nm. A non-fluorescent product is formed upon oxidation. In cationic cethyl trimethyl-1-ammonium chloride (CTAC) and zwitterionic 3-(N-hexadecyl-N,N-dimethylammonium) propane sulfonate (HPS) micellar solutions the same results are observed: three, well-defined, isosbestic points in the optical spectra suggest the transformation between two species. In anionic micellar sodium dodecylsulfate solution (SDS), the appearance of a new band centered around 506 nm was observed, associated to a solution color change from the usual yellow to deep blue/violet, characteristic of a radical species associated to the one-electron oxidation of DIP to its cation radical (DIP*+), observed previously in electrochemical oxidation. Experiments of radical decay kinetics monitoring the absorbance change at 506 nm were performed and analyzed in the frame of a kinetic model taking into account the species both in homogeneous and micellar media. The reaction medium is composed of bulk solution, SDS micelle/solution interface and enzyme catalytic site(s). The variation of DIP*+ concentration was analyzed assuming: (1) synthesis of DIP*+ by HRP through one-electron oxidation; (2) decomposition of DIP*+ by further one-electron oxidation; (3) direct two-electron oxidation of DIP by HRP; (4) bimolecular DIP*+ disproportionation. The main results of the analysis are as follows: (1) kinetic data can be divided in two phases, an HRP active phase and another phase which proceeds in the absence of enzyme activity due to consumption of all H2O2; (2) the reactions of DIP*+ formation, DIP*+ decomposition and DIP two-electron oxidation are HRP concentration dependent; (3) since DIP*+ formation constant seems to be overestimated, it is proposed that two-electron oxidation is another source of DIP*+, through the comproportionation reaction. Evidences for this reaction were also observed previously in electrochemical experiments; and (4) the kinetic analysis provides evidences that the bimolecular reaction of DIP*+ takes place mainly in the absence of active HRP and in this phase the combination of, at least, two second-order kinetic processes is needed to model the experimental data. Our data suggest that HRP oxidizes DIP in general by a two-electron process or that the cation radical is very unstable so that the one-electron process is only detected in the presence of anionic surfactant, which stabilizes significantly the DIP*+ intermediate.  相似文献   

2.
The alkyldiazenes RN = NH (R = CH3 or C2H5) react with reduced microsomal cytochrome P450 leading to complexes exhibiting a Soret peak at 446 nm. Upon oxidation of the [cytochrome P450-Fe(II)(CH3N = NH)] complex with limited amounts of dioxygen, a new complex characterized by a Soret peak at 486 nm is formed. The latter complex was also formed upon slow reaction of methyldiazene with microsomal cytochrome P450-Fe(III) or in situ oxidation of methylhydrazine by limited amounts of O2 or ferricyanide. This complex is rapidly destroyed by O2 or ferricyanide in excess and more slowly by excess dithionite in the presence of CO. Reactions of ethyldiazene or benzyldiazene with cytochrome P450-Fe(III) afforded similar complexes characterized by Soret peaks around 480 nm. These results, when compared to those recently described on reactions of monosubstituted hydrazines RNHNH2 and diazenes RN = NH with hemoglobin and iron-porphyrins, are consistent with a [cytochrome P450-Fe(II)(RN = NH)] structure for the 446-nm-absorbing complexes and a sigma-alkyl cytochrome P450-Fe(III)-R structure for the complexes characterized by a Soret peak around 480 nm. They also suggest a sigma-cytochrome P450-Fe(III)-Ph structure for the complex derived from phenylhydrazine oxidation, recently described in the literature. Finally, they provide the first evidence that cytochrome P450-Fe(III)-R complexes are formed upon microsomal oxidation of alkyl or phenylhydrazines.  相似文献   

3.
P Joliot  A Joliot 《Biochemistry》1999,38(34):11130-11136
Electron transfer within PS I reaction centers has been analyzed in vivo in a mutant of Chlorella sorokiniana which lacks most of the PS II and of the peripheric antennae, using a new spectrophotometric technique with a time resolution of approximately 5 ns. Absorption changes associated with the oxidation of semiphylloquinone (acceptor A(1)(-)) have been characterized in the 371-545 nm spectral range. The oxidation of A(1)(-) and the reduction of an iron-sulfur cluster (F(X), F(A)F(B)) is monitored by an absorption decrease at 377 nm (semiphylloquinone absorption band) and by the decrease of two positive absorption bands around 480 and 515 nm, respectively, very likely associated with a local electrochromic shift induced by A(1)(-) on a carotenoid molecule localized in its vicinity. A(1)(-) undergoes a two-phase oxidation of about equal amplitude with half-times of approximately 18 and approximately 160 ns, respectively. Two hypotheses are proposed to interpret these data: (1) Photosystem I reaction centers are present under two conformational states which differ by the reoxidation rate of A(1)(-). (2) The two phylloquinones corresponding to the two branches of the PS I heterodimer are involved in the electron transfer. The similar amplitude of the two phases implies that the rates of electron transfer from P700 to each of the phylloquinones are about equal. The two different rate constants measured for A(1)(-) oxidation suggests some asymmetry in the relative position of the two phylloquinones with respect to F(X).  相似文献   

4.
Transient spectra and kinetic data of Tiron (1,2-dihydroxybenzene-3,5-disulphonic acid) are reported, obtained after pulse-radiolytic oxidation by hydroxyl radicals (.OH), superoxide anions (O-2) or a combination of both oxygen radicals. The rate constant with .OH radicals was determined at 1.0.10(9) M-1.s-1. Contrary to a previous report (Greenstock, C.L. and Miller, R.W. (1975) Biochim. Biophys. Acta 396, 11--16), the rate constant with O-2 of 1.0.10(7) M-1.s-1 is lower by one order of magnitude; also the semiquinone absorbs at 300 nm rather than at 400 nm. The ratio of the rate constants with .OH and O-2 of 100 again demonstrates that any oxidation reaction by the latter radical is unspecific due to the more efficient reaction of .OH radicals, leading to the same products with catechol compounds.  相似文献   

5.
The water-oxidizing complex of Photosystem II is an important target of ultraviolet-B (280-320 nm) radiation, but the mechanistic background of the UV-B induced damage is not well understood. Here we studied the UV-B sensitivity of Photosystem II in different oxidation states, called S-states of the water-oxidizing complex. Photosystem II centers of isolated spinach thylakoids were synchronized to different distributions of the S(0), S(1), S(2) and S(3) states by using packages of visible light flashes and were exposed to UV-B flashes from an excimer laser (lambda=308 nm). The loss of oxygen evolving activity showed that the extent of UV-B damage is S-state-dependent. Analysis of the data obtained from different synchronizing flash protocols indicated that the UV-sensitivity of Photosystem II is significantly higher in the S(3) and S(2) states than in the S(1) and S(0) states. The data are discussed in terms of a model where UV-B-induced inhibition of water oxidation is caused either by direct absorption within the catalytic manganese cluster or by damaging intermediates of the water oxidation process.  相似文献   

6.
Yang C  Gu ZW  Yang M  Lin SN  Siuzdak G  Smith CV 《Biochemistry》1999,38(48):15903-15908
Oxidative modifications of low-density lipoproteins (LDL) may contribute to the pathogenesis of atherosclerosis. Although the oxidation products of the lipid components of LDL have been studied extensively, less is known about the oxidation products of the apoprotein, apolipoprotein B-100. To identify the specific oxidative modifications, we oxidized LDL in the presence of Cu(2+), treated with DNPH, precipitated and delipidated the protein, digested the protein with trypsin, and analyzed the peptides by high-performance liquid chromatography. We isolated nine peptides that exhibited measurable absorbance at 365 nm, which is characteristic of hydrazones derived from DNPH and is not observed in peptides derived from unoxidized LDL. Unexpectedly, we obtained the same peptides with absorbance at 365 nm in Cu(2+)-oxidized LDL not treated with DNPH. N-terminal sequence analyses and mass spectrometry indicated that the peptides isolated from the Cu(2+)-oxidized LDL all contained kynurenine residues in place of Trp residues found in the native apoprotein. The product profile we observed in Cu(2+)-oxidized LDL was remarkably different from the profiles observed in LDL oxidized by HOCl or myeloperoxidase in vitro, and the preferential oxidation of Trp to kynurenine in Cu(2+)-catalyzed oxidation of LDL contrasts with the products observed following oxidation of LDL with HOCl or myeloperoxidase. Our studies to date support the working hypothesis that the specific products of protein oxidation are sufficiently distinct to be developed as biomarkers of proposed mechanisms of oxidation of LDL and biological molecules in other toxicities and diseases.  相似文献   

7.
Biosynthetic and model in vitro studies have shown that pheomelanins, the distinctive pigments of red human hair, arise by oxidative cyclization of cysteinyldopas mainly 5-S-cysteinyldopa (1) via a critical o-quinonimine intermediate, which rearranges to unstable 1,4-benzothiazines. To get new evidence for these labile species, fast time resolution pulse radiolytic oxidation by dibromide radical anion of a suitable precursor, the dihydro-1,4-benzothiazine-3-carboxylic acid 7 was performed in comparison with that of 1. In the case of 7, dibromide radical anion oxidation leads over a few microseconds (k = 2.1 x 10(9) M(-1) s(-1)) to a phenoxyl radical (lambda(max) 330 nm, epsilon = 6300 M(-1) cm(-1)) which within tens of milliseconds gives rise with second-order kinetics (2k = 2.7 x 10(7) M(-1) s(-1)) to a species exhibiting an absorption maximum at 540 nm (epsilon = 2200 M(-1) cm(-1)). This was formulated as the o-quinonimine 3 arising from disproportionation of the initial radical. The quinonimine chromophore is converted over hundreds of milliseconds (k = 6.0 s(-1)) to a broad maximum at around 330 nm interpreted as due to a 1,4-benzothiazine or a mixture of 1,4-benzothiazines, which as expected are unstable and subsequently decay over a few seconds (k = 0.5 s(-1)). Interestingly, the quinonimine is observed as a labile intermediate also in the alternative reaction route examined, involving cyclization of the o-quinone (lambda(max) 390 nm, epsilon = 6900 M(-1) cm(-1)) arising by disproportionation (2k = 1.7 x 10(8) M(-1) s(-1)) of an o-semiquinone (lambda(max) 320 nm, epsilon = 4700 M(-1) cm(-1)) directly generated by dibromide radical anion oxidation of 1. Structural formulation of the 540 nm species as an o-quinonimine was further supported by rapid scanning diode array spectrophotometric monitoring of the ferricyanide oxidation of a series of model dihydrobenzothiazines.  相似文献   

8.
Bacon Ke  Thomas H. Chaney 《BBA》1971,226(2):341-353
Triton treatment of chromatophores of carotenoid-deficient Chromatium followed by density-gradient centrifugation led to a separation into three subchromatophore fractions. Unlike the case with chromatophores of regular Chromatium, Triton releases about 1/3 of the total bulk bacteriochlorophyll into one fraction (designated G, for green) whose major absorption-band maximum is at 780 nm. One fraction (H, for heavy) absorbs at 805 and 885 nm, with an absorbance ratio A885 nm/A805 nm between 1.5 and 2; another fraction (L, for light) absorbs at 805 nm and has a shoulder at 825 nm. The absorption and fluorescence emission spectra of the three fractions at room temperature and 77°K indicate that the different bacteriochlorophyll forms are efficiently separated by Triton treatment.

The reaction center P890 is concentrated exclusively in the H-fraction, at a level of 5–7% of the bulk bacteriochlorophyll. The solubilized bacteriochlorophyll absorbing at 780 nm can be totally and irreversibly bleached by 5 mM ferricyanide. The other bacteriochlorophyll forms in the H- and L-fractions are also irreversibly bleached by ferricyanide to variable extents. P890 is the only component that can be re-reduced by ascorbate after ferricyanide oxidation. The P890 content estimated by reversible chemical bleaching agrees well with that obtained by reversible light bleaching. The different bacteriochlorophyll forms, with the exception of the 780-nm absorbing form, are relatively stable toward light bleaching. Again, only P890 is reversibly bleached by light.

Cytochromes-555 and -553 are distributed in both the H-and L-fractions, but not in the solubilized-bacteriochlorophyll G-fraction. However, only cytochromes in the H-fraction which contains all of the P890 can undergo coupled oxidation. Excitation with 20-nsec ruby-laser pulses shows that cytochrome-555 can be oxidized in 2–3 μsec by photooxidized P890, indicating that necessary conformation for rapid electron transport is retained in the subchromatophore particles.

The data on fractionation and redox reactions obtained here, together with direct kinetic measurements recently reported in the literature lend further support to the view that oxidation of these two cytochromes is mediated by the same reaction center, P890.  相似文献   


9.
Argyrou A  Blanchard JS  Palfey BA 《Biochemistry》2002,41(49):14580-14590
Lipoamide dehydrogenase catalyses the NAD(+)-dependent oxidation of the dihydrolipoyl cofactors that are covalently attached to the acyltransferase components of the pyruvate dehydrogenase, alpha-ketoglutarate dehydrogenase, and glycine reductase multienzyme complexes. It contains a tightly, but noncovalently, bound FAD and a redox-active disulfide, which cycle between the oxidized and reduced forms during catalysis. The mechanism of reduction of the Mycobacterium tuberculosis lipoamide dehydrogenase by NADH and [4S-(2)H]-NADH was studied anaerobically at 4 degrees C and pH 7.5 by stopped-flow spectrophotometry. Three phases of enzyme reduction were observed. The first phase, characterized by a decrease in absorbance at 400-500 nm and an increase in absorbance at 550-700 nm, was fast (k(for) = 1260 s(-)(1), k(rev) = 590 s(-)(1)) and represents the formation of FADH(2).NAD(+), an intermediate that has never been observed before in any wild-type lipoamide dehydrogenase. A primary deuterium kinetic isotope effect [(D)(k(for) + k(rev)) approximately 4.2] was observed on this phase. The second phase, characterized by regain of the absorbance at 400-500 nm, loss of the 550-700 nm absorbance, and gain of 500-550 nm absorbance, was slower (k(obs) = 200 s(-)(1)). This phase represents the intramolecular transfer of electrons from FADH(2) to the redox-active disulfide to generate the anaerobically stable two-electron reduced enzyme, EH(2). The third phase, characterized by a decrease in absorbance at 400-550 nm, represents the formation of the four-electron reduced form of the enzyme, EH(4). The observed rate constant for this phase showed a decreasing NADH concentration dependence, and results from the slow (k(for) = 57 s(-)(1), k(rev) = 128 s(-)(1)) isomerization of EH(2) or slow release of NAD(+) before rapid NADH binding and reaction to form EH(4). The mechanism of oxidation of EH(2) by NAD(+) was also investigated under the same conditions. The 530 nm charge-transfer absorbance of EH(2) shifted to 600 nm upon NAD(+) binding in the dead time of mixing of the stopped-flow instrument and represents formation of the EH(2).NAD(+) complex. This was followed by two phases. The first phase (k(obs) = 750 s(-)(1)), characterized by a small decrease in absorbance at 435 and 458 nm, probably represents limited accumulation of FADH(2).NAD(+). The second phase was characterized by an increase in absorbance at 435 and 458 nm and a decrease in absorbance at 530 and 670 nm. The observed rate constant that describes this phase of approximately 115 s(-)(1) probably represents the overall rate of formation of E(ox) and NADH from EH(2) and NAD(+), and is largely determined by the slower rates of the coupled sequence of reactions preceding flavin oxidation.  相似文献   

10.
EDTA not only blocks the horseradish peroxidase (HRP)-catalyzed iodide oxidation to I-3 but also causes an enzymatic conversion of oxidized iodine species to iodide (Banerjee, R. K., De, S. K., Bose, A. K., and Datta, A. G. (1986) J. Biol. Chem. 261, 10592-10597). The EDTA effect on both of these reactions can be withdrawn with a higher concentration of iodide and not with H2O2. Spectral studies indicate a possible interaction of EDTA with HRP as evidenced by the formation of modified compound 1 with H2O2 at 416 nm instead of 412 nm in the absence of EDTA. EDTA causes a hypochromic effect on HRP at 402 nm which undergoes the bathochromic red shift to 416 nm by H2O2. The addition of iodide to the 416 nm complex causes the reappearance of the Soret band of HRP at 402 nm. Among various EDTA analogues tested, N-N-N'-N'-tetramethylethylenediamine (TEMED) is 80% as effective as EDTA in the conversion of I-3 to iodide and produces a spectral shift of HRP similar to EDTA. Interaction of EDTA with HRP is further indicated by the hyperchromic effect of HRP and H2O2 on the absorption of EDTA at 212 nm. The addition of oxidized iodine species produces a new peak at 230 nm due to formation of iodide. EDTA at a higher concentration can effectively displace radioiodide specifically bound to HRP indicating its interaction at the iodide-binding site. The enzyme, after radioiodide displacement with EDTA, shows a characteristic absorption maximum at 416 nm on the addition of H2O2, indicating that EDTA is bound with the enzyme. Both positive and negative circular dichroism spectra of HRP and the HRP.H2O2 complex, characteristic of heme absorption, are altered by EDTA, suggesting an EDTA-induced conformational change at or near the heme region. This is associated with a change of affinity of heme toward H2O2 and azide. It is postulated that EDTA interacts at the iodide-binding site of the HRP inducing a new conformation that blocks iodide oxidation but is suitable to convert iodine to iodide by a redox reaction with H2O2.  相似文献   

11.
Phagocytes produce large quantities of reactive oxygen species for pathogen killing; however, the kinetics and amplitude of ROS production on the level of individual phagosomes are poorly understood. This is mainly due to the lack of appropriate methods for quantitative ROS detection with microscopic resolution. We covalently attached the ROS-sensitive dye dichlorodihydrofluorescein (DCFH(2)) to yeast particles and investigated their fluorescence due to oxidation in vitro and in live phagocytes. In vitro, the dye was oxidized by H(2)O(2) plus horseradish peroxidase but also by HOCl. The latter produced a previously unrecognized oxidation product with red-shifted excitation and emission spectra and a characteristic difference in the shape of the excitation spectrum near 480 nm. Millimolar HOCl bleached the DCFH(2) oxidation products. Inside phagosomes, DCFH(2)-labeled yeast were oxidized for several minutes in a strictly NADPH oxidase-dependent manner as shown by video microscopy. Inhibition of the NADPH oxidase rapidly stopped the fluorescence increase of the particles. At least two characteristic kinetics of oxidation were distinguished and the variability of DCFH(2) oxidation in phagosomes was much larger than the variability upon oxidation in vitro. We conclude that DCFH(2)-yeast is a valuable tool to investigate the kinetics and amplitude of ROS production in individual phagosomes.  相似文献   

12.
Pseudomonas AM1, Hyphomicrobium X and Pseudomonas MS all contain cytochrome a/a(3) and a b-type cytochrome able to react with CO. Pseudomonas AM1 and Hyphomicrobium X also have a CO-binding cytochrome c. The purified cytochrome c (redox potential 0.26V) of Pseudomonas AM1 was not susceptible to oxidation by molecular oxygen. CO reacted slowly with the reduced form giving a CO difference spectrum with a peak at 412nm and troughs at 420nm and 550nm. Similar results were obtained with the cytochrome c of Hyphomicrobium (aerobically grown or anaerobically grown with nitrate) and with that of Pseudomonas extorquens. The results given in the present paper are incompatible with an oxygenase or oxidase function for the soluble cytochrome c of methylotrophs. Studies with whole cells of Pseudomonas AM1 and a cytochrome c-deficient mutant have demonstrated that cytochrome b (redox potential 0.009V) is the first cytochrome in the electron-transport chain for oxidation of all substrates except methanol (and ethanol) whose oxidation does not involve this cytochrome. All substrates are usually oxidized by way of cytochrome c and cytochrome oxidase (cytochrome a/a(3)), but there is an alternative route for the reduction of cytochrome a/a(3) in the mutant lacking cytochrome c. Results of experiments on cyanide inhibition of respiration and cytochrome oxidation support the suggestion that the susceptibility of cytochrome b to oxidation by molecular oxygen (reflected in its ability to react with CO) is probably irrelevant to the normal physiology of Pseudomonas AM1.  相似文献   

13.
Rutin (3',4',5,7-tetrahydroxyflavone-3-rutinoside) was oxidized by a horseradish peroxidase-H2O2 system to an ascorbate-reducible product which had an absorption maximum at about 290 nm and a shoulder at about 440 nm at pH 4. At pH 7.8, ascorbate-reducible compounds and sodium hydrosulfite-reducible and -nonreducible compounds were formed by the oxidation. The ascorbate-reducible compounds consisted of at least two components, the absorption bands of which were at 460-480 nm and about 620 nm. The sodium hydrosulfite-reducible compounds also consisted of two components, and one of the components which had an absorption maximum at about 480 nm seems to be formed from the ascorbate-reducible component of an absorption maximum at the blue region by a nonenzymatic reaction. A mixture of oxidized products of rutin formed by tert-butyl hydroperoxide-dependent oxidation was similar to that formed by the enzymatic reaction. It is discussed that the 3'- and 4'-OH groups of rutin were oxidized by the horseradish peroxidase-H2O2 system and that the oxidized product which could be reduced by ascorbate is an o-quinone derivative.  相似文献   

14.
The mechanisms of oxidation of low-density lipoproteins (LDLs) are not well defined, but epidemiological and experimental studies suggest that iron-catalyzed processes may contribute to atherogenesis. The aim of this study was to test the hypothesis that iron-catalyzed oxidations of LDLs in vitro produce diagnostic biomarkers of oxidation of the apolipoprotein that could be applied to studies in vivo. LDLs were oxidized in the presence of Fe2+, EDTA, and ascorbic acid for up to 40 h. Following delipidation and trypsin digestion, the peptides were separated by HPLC, with four peaks detected at 365 nm, whereas none were observed in peptides from unoxidized LDLs. The peptides were identified by MALDI-QTOF mass spectrometry as IVQILP(W+4) EQNEQVK, IYSL(W+4)EHSTK, FEGLQE(W+4)EGK, and YH(W+4)EHTGLTLR, with (W+4) rather than the W residues of the unoxidized protein. The mass gains (+4 increase in m/z in tryptophan, W) and absorbance at 365 nm indicate kynurenines, which were trypsin-releasable peptides that are on the surface of LDL particles. All four peptides thus characterized share the sequence of WE. The preferential oxidation of W residues in WE sequences suggest contributions from the C-proximate glutamate residues in chelation of the iron species, thereby influencing site selectivities of oxidation. These kynurenine-containing peptides might serve as biomarkers of iron-mediated oxidations in vivo.  相似文献   

15.
1. Cytochrome c2+ increases the rate at which cytochrome oxidase (EC 1.9.3.1) gamma max428nm) converts to its conformational isomer (gamma max 418-423 nm) but cytochrome c3+ has little effect on the conversion rate. 2. Interactions between reduced cytochrome oxidase and cytochrome c were studied in the absence of electron flow using anaerobic Sephadex columns. 3. Oxidase that is reduced by cytochrome c2+ or other reductant forms the 418-to 423-nm isomer if its last contact, before oxidation, is with cytochrome c3+. If the reduced oxidase contacts cytochrome c2+, before oxidation, the 428-nm oxidase forms.  相似文献   

16.
In the presence of 0.1-5 muM N-methylphenazonium methosulphate approx. 50-70% oxidation of cytochrome b-559 can be induced by far-red light. The oxidation is best observed with long wavelength far-red light (732 nm) of moderate intensities (approx. 10(4) ergs/cm2 per s) and is reversed by subsequent illumination with red light. Concentrations of N-methylphenazonium methosulphate above 5 muM are inhibitory probably due to cyclic electron flow. The far-red oxidation is inhibited by low concentrations of the plastoquinone antagonist 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone, while 3-(3,4-dichlorophenyl)-1,1-dimethylurea inhibits red light reduction and increases the amplitude of far-red oxidation. The effect of N-methylphenazonium methosulphate is mimicked by N-methyl-phenazonium ethosulphate, but not by pyocyanine or diaminodurene. Low concentrations (2-3 muM) of N-methylphenazonium methosulphate also stimulate a 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone-inhibitable red light reduction of cytochrome f.  相似文献   

17.
Horseradish peroxidase (HRP) catalyzes the polymerization of free heme (beta-hematin formation) through its oxidation. Heme when added to HRP compound II (FeIV=O) causes spectral shift from 417 nm (Compound II) to 402 nm (native, FeIII) indicating that heme may be oxidized via one-electron transfer. Direct evidence for one-electron oxidation of heme by HRP intermediates is provided by the appearance of an E.s.r signal of a 5,5-dimethyl-1-pyrroline N-oxide (spin trap)-heme radical adduct (a1H=14.75 G, a2H=4.0 G) in E.s.r studies. Heme-polymerization by HRP is inhibited by spin trap indicating that one-electron oxidation product of heme ultimately leads to the formation of heme-polymer. HRP, when incubated with diethyl pyrocarbonate (DEPC), a histidine specific reagent, shows concentration dependent loss of heme-polymerization indicating the role of histidine residues in the process. We suggest that HRP catalyzes the formation of heme-polymer through one-electron oxidation of free heme.  相似文献   

18.
Electrolysis or horseradish peroxidase (HRP)-catalyzed oxidation of tyrosine and bityrosine in aqueous solution at pH 7.4 resulted in light emission in the visible region. Electrolysis of tyrosine emitted light which peaked at 490 nm and was almost completely quenched by superoxide dismutase (SOD), while emission by bityrosine peaked at 530 nm. In the HRP-H(2)O(2)-tyrosine system the oxidation-reduction of tyrosine emitted light with two prominent peaks, 490 and 530 nm, and was not quenched by SOD. The phenoxyl neutral radical of the tyrosine in HRP-H(2)O(2)-tyrosine system was detected by electron spin resonance (ESR) spectrometry using tert-nitrosobutane as a spin trap; the spin adduct was found to adhere to the HRP molecule during the enzymatic reaction. Further, bityrosine was detected in the HRP-H(2)O(2)-tyrosine reaction system. Changes in absorption spectra of HRP and chemiluminescence intensities during HRP-catalyzed oxidation of tyrosine suggest that for photon emission compound III is a candidate superoxide donor to the phenoxyl cation radical of tyrosine on the enzyme molecule. The luminescence observed in this study might be originated from at least two exciplexes involved with the tyrosine cation radical (Tyr(*+)) and the bityrosine cation radical (BT(*+))  相似文献   

19.
Degradation of the reduced pyridine nucleotides NMNH and NADH by HOCl involves two distinct stages: a fast reaction, k = 4.2 x 10(5) M(-1) s(-1), leads to generation of stable pyridine products (Py/Cl) with a strong absorption band at 275 nm (epsilon = 12.4 x 10(3) M(-1) cm(-1) in the case of NMNH); secondarily, a subsequent reaction of HOCl, k = 3.9 x 10(3) M(-1) s(-1), leads to a complete loss of the aromatic absorption band of the pyridine ring. HOBr and HOI(I(2)) react similarly. Apparent rate constants of the primary reactions of HOX species with NMNH at pH 7.2 increase in the order HOCl (3 x 10(5) M(-1) s(-1)) < HOBr( approximately 4 x 10(6) M(-1) s(-1)) < HOI(I(2))( approximately 6.5 x 10(7) M(-1) s(-1)). HOBr reacts fast also with the primary product Py/Br, k approximately 9 x 10(5) M(-1) s(-1), while the reactions of HOI and I(2) with Py/I are slower, approximately 1.4 x 10(3) M(-1) s(-1) and >6 x 10(3) M(-1) s(-1), respectively. Halogenation of the amide group of NMN(+) by HOX species is many orders of magnitude slower than oxidation of NMNH. Taurine inhibits HOCl-induced oxidation of NADH, but HOBr-induced oxidation is not inhibited because the taurine monobromamine rapidly oxidizes NADH, and oxidation by HOI(I(2)) is not inhibited because taurine is inert toward HOI(I(2)). Also sulfur compounds (GSH, GSSG, and methionine) are less efficient in protecting NADH against oxidation by HOBr and HOI(I(2)) than against oxidation by HOCl. The results suggest that reactions of HOBr and HOI(I(2)) in a cellular environment are much more selectively directed toward irreversible oxidation of NADH than reactions of HOCl. It is noteworthy that the rather inert N-chloramines react with iodide to generate HOI(I(2)), i.e., the most reactive and selective oxidant of reduced pyridine nucleotides. NMR investigations show that the primary stable products of the reaction between NMNH and HOCl are various isomeric chlorohydrins originating from a nonstereospecific electrophilic addition of HOCl to the C5&dbond;C6 double bond of the pyridine ring. The primary products (Py/X) of NMNH all exhibit similar absorption bands around 275 nm and are hence likely to result from analogous addition of HOX to the C5&dbond;C6 bond of the pyridine ring. Since the Py/X species are stable and inert toward endogeneous reductants like ascorbate and GSH, they may generally be useful markers for assessing the contribution of hypohalous acids to inflammatory injury.  相似文献   

20.
Modification of low density lipoproteins (LDL) by oxidation has been shown to permit recognition by the acetyl-LDL receptor of macrophages. The extensive oxidation of LDL that is required before interaction occurs with this receptor produces major alterations in both the lipid and protein components of LDL. Several chemical modifications of LDL also lead to recognition by this receptor; all of these involve derivatization of lysine residues of apolipoprotein B by adducts that neutralize the positively charged epsilon-amino group. The present studies show that oxidation also results in derivatization of LDL lysine residues. Analysis of amino acid composition indicated that 32% of lysine residues were modified after oxidation of LDL by exposure to 5 microM CuSO4 for 20 h. About one-half of the derivatized lysines were labile under the conditions of acid hydrolysis. Fluorescence of LDL protein was greatly increased by oxidation, with excitation maximum at 350 nm and emission maximum at 433 nm. When LDL containing phosphatidylcholine with isotopically labeled arachidonic acid in the sn-2 position was oxidized, there was a 5-fold increase in radioactivity bound to protein compared to nonoxidized LDL or oxidized LDL labeled with 2-[1-14C]palmitoyl phosphatidylcholine. Prior methylation of LDL prevented the rapid uptake and degradation by macrophages that normally accompanies oxidation. These findings suggest that oxidation of LDL is accompanied by derivatization of lysine epsilon-amino groups by lipid products and that these adducts may be important in the interaction of oxidized LDL with the acetyl-LDL receptor.  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