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 15 毫秒
1.
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   

2.
Rat liver microsomes catalyze a vanadate-stimulated oxidation of NAD(P)H, which is augmented by paraquat and suppressed by superoxide dismutase, but not by catalase. NADPH oxidation was a linear function of the concentration of microsomes in the absence of vanadate, but was a saturating function in the presence of vanadate. Microsomes did not catalyze a vanadate-stimulated oxidation of reduced nicotinamide mononucleotide (NMNH), but gained this ability when NADPH was also present. When the concentration of NMNH was much greater than that of NADPH a minimal average chain length could be calculated from 1/2 the ratio of NMNH oxidized per NADPH added. The term chain length, as used here, signifies the number of molecules of NMNH oxidized per initiating event. Chain length could be increased by increasing [vanadate] and [NMNH] and by decreasing pH. Chain lengths in excess of 30 could easily be achieved. The Km for NADPH, arrived at from saturation of its ability to trigger NMNH oxidation by microsomes in the presence of vanadate, was 1.5 microM. Microsomes or the outer mitochondrial membrane was able to catalyze the vanadate-stimulated oxidation of NADH or NADPH but only the oxidation of NADPH was accelerated by paraquat. The inner mitochondrial membrane was able to cause the vanadate-stimulated oxidation of NAD(P)H and in this case paraquat stimulated the oxidation of both pyridine coenzymes. Our results indicate that vanadate stimulation of NAD(P)H oxidation by biomembranes is a consequence of vanadate stimulation of NAD(P)H or NMNH oxidation by O-2, rather than being due to the existence of vanadate-stimulated NAD(P)H oxidases or dehydrogenases.  相似文献   

3.
Reactive oxygen species produced by activated neutrophils and monocytes are thought to be involved in mediating the loss of collagen and other matrix proteins at sites of inflammation. To evaluate their potential to oxidize the pyridinoline (Pyd) cross-links found in collagen types I and II, we reacted hydrogen peroxide (H(2)O(2)), hypochlorous acid/hypochlorite (HOCl/OCl(-)), and singlet oxygen (O(2)((1)delta g)) with the Pyd substitutes, pyridoxamine dihydrochloride and vitamin B(6), which share the same chemical structure and spectral properties of Pyd cross-links. Neither H(2)O(2) (125-500 microm) nor O(2)((1)delta g) (10-25 microm) significantly changed the spectral properties of pyridoxamine or vitamin B(6). Reaction of HOCl/OCl(-) (12.5-50 microm) with pyridoxamine at pH 7.2 resulted in a concentration-dependent appearance of two new absorbance peaks and a decrease in fluorescence at 400 nm (excitation 325 nm). The new absorbance peaks correlated with the formation of an N-chloramine and the product of its subsequent reaction with pyridoxamine. In contrast, the extent to which HOCl reacted with vitamin B(6), which lacks a primary amine group, was variable at this pH. At lysosomal pH 5.5, Cl(2)/HOCl/OCl(-) reacted with both pyridoxamine and vitamin B(6). Four of the chlorinated products of this reaction were identified by gas chromatography-mass spectrometry and included 3-chloropyridinium, an aldehyde, and several chlorinated products with disrupted rings. To evaluate the effects of Cl(2)/HOCl/OCl(-) on Pyd cross-links in collagen, we exposed bone collagen type I and articular cartilage type II to HOCl. Treatment of either collagen type with HOCl at pH 5. 0 or 7.2 resulted in the oxidation of amine groups and, for collagen type II, the specific decrease in Pyd cross-link fluorescence, suggesting that during inflammation both oxidations may be used by neutrophils and monocytes to promote the loss of matrix integrity.  相似文献   

4.
R S Ehrlich  R F Colman 《Biochemistry》1985,24(20):5378-5387
The binding of coenzymes, NADP+ and NADPH, and coenzyme fragments, 2'-phosphoadenosine 5'-(diphosphoribose), adenosine 2',5'-bisphosphate, and 2'-AMP, to pig heart NADP+-dependent isocitrate dehydrogenase has been studied by proton NMR. Transferred nuclear Overhauser enhancement (NOE) between the nicotinamide 1'-ribose proton and the 2-nicotinamide ring proton indicates that the nicotinamide-ribose bond assumes an anti conformation. For all nucleotides, a nuclear Overhauser effect between the adenine 1'-ribose proton and 8-adenine ring proton is observed, suggesting a predominantly syn adenine--ribose bond conformation for the enzyme-bound nucleotides. Transferred NOE between the protons at A2 and N6 is observed for NADPH (but not NADP+), implying proximity between adenine and nicotinamide rings in a folded enzyme-bound form of NADPH. Line-width measurements on the resonances of free nucleotides exchanging with bound species indicate dissociation rates ranging from less than 7 s-1 for NADPH to approximately 1600 s-1 for adenosine 2',5'-bisphosphate. Substrate, magnesium isocitrate, increases the dissociation rate for NADPH about 10-fold but decreases the corresponding rate for phosphoadenosine diphosphoribose and adenosine 2',5'-bisphosphate about 10-fold. These effects are consistent with changes in equilibrium dissociation constants measured under similar conditions. The 1H NMR spectrum of isocitrate dehydrogenase at pH 7.5 has three narrow peaks between delta 7.85 and 7.69 that shift with changes in pH and hence arise from C-4 protons of histidines. One of those, with pK = 5.35, is perturbed by NADP+ and NADPH but not by nucleotide fragments, indicating that this histidine is in the region of the nicotinamide binding site. Observation of nuclear Overhauser effects arising from selective irradiation at delta 7.55 indicates proximity of either a nontitrating histidine or an aromatic residue to the adenine ring of all nucleotides. In addition, selective irradiation of the methyl region of the enzyme spectrum demonstrates that the adenine ring is close to methyl side chains. The substrate magnesium isocitrate produces no observable differences in these protein--nucleotide interactions. The alterations in enzyme--nucleotide conformation that result in changes in affinity in the presence of substrate must involve either small shifts in the positions of amino acid side chains or changes in groups not visible in the proton NMR spectrum.  相似文献   

5.
J.B. Hoek  L. Ernster  E.J. De Haan  J.M. Tager 《BBA》1974,333(3):546-559
1. The kinetics of oxidation of intramitochondrial reduced nicotinamide nucleotides by -oxoglutarate plus ammonia in intact rat-liver mitochondria have been reinvestigated. It is demonstrated that the preferential oxidation of NADPH observed on addition of ammonia to mitochondria, preincubated under energized conditions in the presence of -oxoglutarate, is due to a transhydrogenation catalysed by glutamate dehydrogenase rather than to an energy-dependent modification of the nicotinamide nucleotide specificity of the enzyme in intact mitochondria.

2. When mitochondria are preincubated at 25 °C under energized conditions in the presence of respiratory inhibitors with the substrates of glutamate dehydrogenase, an oxidation of NADPH, but not of NADH, is brought about by decreasing the reaction temperature. Both the rate of NADPH oxidation and the final steady-state mass-action ratio of nicotinamide nucleotides are dependent on the concentration of ammonia and on the final reaction temperature. A similar effect is observed when rhein is added to the reaction medium at 25 °C in order to inhibit the energy-linked transhydrogenase reaction.

3. In the presence of the substrates of glutamate dehydrogenase, intact ratliver mitochondria catalyse an ATPase reaction due to the simultaneous activity of the energy-linked transhydrogenase and the non-energy-linked transhydrogenation catalysed by glutamate dehydrogenase.

4. These findings are discussed in relation to the nicotinamide nucleotide specificity of glutamate dehydrogenase and to a possible compartmentation of nicotinamide nucleotides in intact rat-liver mitochondria.  相似文献   


6.
Irreversible oxidation of reduced nicotinamide nucleotides by neutrophil-derived halogen oxidants (HOCl, chloramines, HOBr, etc.) is likely to be a highly lethal process, because of the essential role of NAD(P)H in important cell functions such as mitochondrial electron transport, and control of the cellular thiol redox state by NADPH-dependent glutathione reductase. Chloramines (chloramine-T, NH(2)Cl, etc.) and N-chloramides (N-chlorinated cyclopeptides) react with NADH to generate the same products as HOCl, i.e., pyridine chlorohydrins, as judged from characteristic changes in the NADH absorption spectrum. Compared with the fast oxidation of NADH by HOCl, k approximately 3 x 10(5) M(-1) s(-1) at pH 7.2, the oxidation by chloramines is about five orders of magnitude slower; that by chloramides is about four orders of magnitude slower. Apparent rate constants for oxidation of NADH by chloramines increase with increasing proton or buffer concentration, consistent with general acid catalysis, but oxidation by chloramides proceeds with pH-independent kinetics. In presence of iodide the oxidation of NADH by chloramines or chloramides is faster by at least two orders of magnitude; this is due to reaction of iodide with the N-halogen to give HOI/I(2), the most reactive and selective oxidant for NADH among HOX species. Quinuclidine derivatives (QN) like 3-chloroquinuclidine and quinine are capable of catalyzing the irreversible degradation of NADH by HOCl and by chloramines; QN(+)Cl, the chain carrier of the catalytic cycle, is even more reactive toward NADH than HOCl/ClO(-) at physiological pH. Oxidation of NADH by NH(2)Br proceeds by fast, but complex, biphasic kinetics. A compilation of rate constants for interactions of reactive halogen species with various substrates is presented and the concept of selective reactivity of N-halogens is discussed.  相似文献   

7.
The NAD(P)H:flavin oxidoreductase from Escherichia coli, named Fre, is a monomer of 26.2 kDa that catalyzes the reduction of free flavins using NADPH or NADH as electron donor. The enzyme does not contain any prosthetic group but accommodates both the reduced pyridine nucleotide and the flavin in a ternary complex prior to oxidoreduction. The specificity of the flavin reductase for the pyridine nucleotide was studied by steady-state kinetics using a variety of NADP analogs. Both the nicotinamide ring and the adenosine part of the substrate molecule have been found to be important for binding to the polypeptide chain. However, in the case of NADPH, the 2'-phosphate group destabilized almost completely the interaction with the adenosine moiety. Moreover, NADPH and NMNH are very good substrates for the flavin reductase, and we have shown that both these molecules bind to the enzyme almost exclusively by the nicotinamide ring. This provides evidence that the flavin reductase exhibits a unique mode for recognition of the reduced pyridine nucleotide. In addition, we have shown that the flavin reductase selectively transfers the pro-R hydrogen from the C-4 position of the nicotinamide ring and is therefore classified as an A-side-specific enzyme.  相似文献   

8.
G F Leanz  G G Hammes 《Biochemistry》1986,25(19):5617-5624
The ionic strength dependence of the second-order rate constant for the association of reduced nicotinamide adenine dinucleotide phosphate (NADPH) and chicken liver fatty acid synthase was determined. This rate constant is 7.2 X 10(7) M-1 s-1 at zero ionic strength and 25 degrees C; the effective charge at the cofactor binding sites is +0.8. The conformations of nicotinamide adenine dinucleotide phosphate (NADP+) and NADPH bound to the beta-ketoacyl and enoyl reductase sites were determined from transferred nuclear Overhauser effect measurements. Covalent modification of the enzyme with pyridoxal 5'-phosphate abolished cofactor binding at the enoyl reductase site; this permitted the cofactor conformations at the beta-ketoacyl and enoyl reductase sites to be distinguished. For NADP+ bound to the enzyme, the conformation of the nicotinamide-ribose bond is anti at the enoyl reductase site and syn at the beta-ketoacyl reductase site; the adenine-ribose bond is anti, and the sugar puckers are C3'-endo. Nicotinamide-adenine base stacking was not detected. Structural models of NADP+ at the beta-ketoacyl and enoyl reductase sites were constructed by using the distances calculated from the observed nuclear Overhauser effects. Because of the overlap of the resonances of several nonaromatic NADPH protons with the resonances of HDO and ribose protons, less extensive structural information was obtained for NADPH bound to the enzyme. However, the conformations of NADPH bound to the two reductases are qualitatively the same as those of NADP+, except that the nicotinamide moiety of NADPH is closer to being fully anti at the enoyl reductase site.  相似文献   

9.
Vandate augments the oxidation of NAD(P)H, but not of NMNH, by rat liver microsomes. Paraquat increases the vanadate effect on NADPH, but not on NADH, oxidation. Substoichiometric levels of NADPH caused the co-oxidation of NADH or NMNH and SOD inhibited in all cases. The ratio of NADH or NMNH co-oxidized per NADPH added allowed estimation of average chain length, which increased as the pH was lowered from 8.0 to 7.1. The initial rate of this co-oxidation of NMNH was a saturating function of the concentration of microsomes, reflecting a decrease in chain length with an increase in number of concomitant reaction chains, and due to increasing radical-radical termination reactions. Mitochondrial outer membranes behaved like the microsomal membranes, but mitochondrial inner membranes catalyzed a rapid oxidation of NADH which could be augmented by vanadate, whose action was enhanced by paraquat and inhibited by antimycin or rotenone. These and related observations support the view that vanadate stimulates NAD(P)H oxidation by biological membranes, not by virtue of interacting with enzymes, but rather by interacting with O-2.  相似文献   

10.
Most studies of protein oxidation have typically focused on the reactivity of single amino acid side chains while ignoring the potential importance of adjacent sequences in directing the reaction pathway. We previously showed that hypochlorous acid (HOCl), a specific product of myeloperoxidase, inactivates matrilysin by modifying adjacent tryptophan and glycine (WG) residues in the catalytic domain. Here, we use model peptides that mimic the region of matrilysin involved in this reaction, VVWGTA, VVWATA, and the library VVWXTA, to determine whether specific sequence motifs are targeted for chlorination or oxygenation by myeloperoxidase. Our results demonstrate that HOCl generated by myeloperoxidase or activated neutrophils converts the peptide VVWGTA to a chlorinated product, WG+32(Cl). Tandem mass spectrometry in concert with high resolution 1H and two-dimensional NMR analysis revealed that the modification required cross-linking of the tryptophan to the amide of glycine followed by chlorination of the indole ring of tryptophan. In contrast, when glycine in the peptide was replaced with alanine, the major products were mono- and dioxygenated tryptophan residues. When the peptide library VVWXTA (where X represents all 20 common amino acids) was exposed to HOCl, only WG produced a high yield of the chloroindolenine derivative. However, when glycine was replaced by other amino acids, oxygenated tryptophan derivatives were the major products. Our observations indicate that WG may represent a specific sequence motif in proteins that is targeted for chlorination by myeloperoxidase.  相似文献   

11.
Hypochlorous acid (HOCl) is formed by the action of the enzyme myeloperoxidase on hydrogen peroxide and chloride ions. It has been shown to be highly bactericidal and cytotoxic by a variety of mechanisms, one of which, may be the modification of DNA. Previously we have demonstrated by GC-MS analysis that exposure of calf thymus DNA to HOCl causes extensive pyrimidine modification, including 5-chlorocytosine formation. Using GC-MS analysis, we now demonstrate the formation of an additional chlorinated base product, 8-Cl adenine. The addition of 50 μM HOCl was sufficient to produce a significant increase in this product. The reaction of HOCl with adenine in calf thymus DNA was shown to be rapid with the reaction complete after 1 min. pH-dependence studies suggest HOCl rather than its conjugate base (OCl-) to be responsible for 8-Cl adenine formation. Other commercially available chlorinated base products, 6-Cl guanine or 2-Cl adenine were not detected. Therefore, 8-Cl adenine might prove a useful biomarker for studying the role of reactive chlorine species (RCS) during inflammatory processes.  相似文献   

12.
M F Carlier  D Pantaloni 《Biochemistry》1976,15(21):4703-4712
The binding of reduced nicotinamide adenine dinucleotide phosphate (NADPH) to nicotinamide adenine dinucleotide phosphate (NADP) dependent isocitrate dehydrogenase from beef liver cytoplasm was studied by several equilibrium techniques (ultracentrifugation, molecular sieving, ultrafiltration, fluorescence). Two binding sites (per dimeric enzyme molecule) were found with slightly different dissociation constants (0.5 and 0.12 muM) and fluorescence yields (7.7 and 6.3). A ternary complex was formed between enzyme, isocitrate, and NADPH, in which NADPH dissociation constant was 5 muM. On the contrary, no binding of NADPH to the enzyme took place in the presence of magnesium isocitrate. Dialysis experiments showed the existence of 1 NADP binding site/dimer, with a dissociation constant of 26 muM. When NADPH was present with the enzyme in the proportion of 1 molecule/dimer, the dissociation constant of NADP was decreased fourfold, reaching a value quantitatively comparable to the Michaelis constant. The kinetics of coenzyme binding was followed using the stopped-flow technique with fluorescence detection. NADPH binding to the enzyme occurred through one fast reaction (k1 = 20 muM-1 s-1). Dissociation of NADPH took place upon NADP binding; however, equilibrium as well as kinetic data were incompatible with a simple competition scheme. Dissociation of NADPH from the enzyme upon magnesium isocitrate binding was preceded by the formation of a transitory ternary complex in which the fluorescence of NADPH was only about 30% of that in the enzyme-NADPH complex. Then interaction between the conenzymes and the involvement of ternary complexes in the catalytic mechanism are discussed in relation with what is known about the regulatory role of the coenzyme (Carlier, M. F., and Pantaloni, D. (1976), Biochemistry, 15, 1761-1766).  相似文献   

13.
We have prepared a selectively deuterated dihydrofolate reductase in which all the aromatic protons except the C(2) protons of tryptophan have been replaced by deuterium and have examined the 1H NMR spectra of its complexes with folate, trimethoprim, methotrexate, NADP+, and NADPH. One of the four Trp C(2)-proton resonance signals (signal P at 3.66 ppm from dioxane) has been assigned to Trp-21 by examining the NMR spectrum of a selectively deuterated N-bromosuccinimide-modified dihydrofolate reductase. This signal is not perturbed by NADPH, indicating that the coenzyme is not binding close to the 2 position of Trp-21. This contrasts markedly with the 19F shift (2.7 ppm) observed for the 19F signal of Trp-21 in the NADPH complex with the 6-fluorotryptophan-labeled enzyme. In fact the crystal structure of the enzyme . methotrexate . NADPH shows that the carboxamide group of the reduced nicotinamide ring is near to the 6 position of Trp-21 but remote from its 2 position. The nonadditivity of the 1H chemical-shift contributions for signals tentatively assigned to Trp-5 and -133 indicates that these residues are influenced by ligand-induced conformational changes.  相似文献   

14.
Myeloperoxidase, a heme enzyme secreted by activated phagocytes, uses H(2)O(2) and Cl(-) to generate the chlorinating intermediate hypochlorous acid (HOCl). This potent cytotoxic oxidant plays a critical role in host defenses against invading pathogens. In this study, we explore the possibility that myeloperoxidase-derived HOCl might oxidize nucleic acids. When we exposed 2'-deoxycytidine to the myeloperoxidase-H(2)O(2)-Cl(-) system, we obtained a single major product that was identified as 5-chloro-2'-deoxycytidine using mass spectrometry, high performance liquid chromatography, UV-visible spectroscopy, and NMR spectroscopy. 5-Chloro-2'-deoxycytidine production by myeloperoxidase required H(2)O(2) and Cl(-), suggesting that HOCl is an intermediate in the reaction. However, reagent HOCl failed to generate 5-chloro-2'-deoxycytidine in the absence of Cl(-). Moreover, chlorination of 2'-deoxycytidine was optimal under acidic conditions in the presence of Cl(-). These results implicate molecular chlorine (Cl(2)), which is in equilibrium with HOCl through a reaction requiring Cl(-) and H(+), in the generation of 5-chloro-2'-deoxycytidine. Activated human neutrophils were able to generate 5-chloro-2'-deoxycytidine. Cellular chlorination was blocked by catalase and heme poisons, consistent with a myeloperoxidase-catalyzed reaction. The myeloperoxidase-H(2)O(2)-Cl(-) system generated similar levels of 5-chlorocytosine in RNA and DNA in vitro. In striking contrast, only cell-associated RNA acquired detectable levels of 5-chlorocytosine when intact Escherichia coli was exposed to the myeloperoxidase system. This observation suggests that oxidizing intermediates generated by myeloperoxidase selectively target intracellular RNA for chlorination. Collectively, these results indicate that Cl(2) derived from HOCl generates 5-chloro-2'-deoxycytidine during the myeloperoxidase-catalyzed oxidation of 2'-deoxycytidine. Phagocytic generation of Cl(2) therefore may constitute one mechanism for oxidizing nucleic acids at sites of inflammation.  相似文献   

15.
Transhydrogenase couples the redox reaction between NADH and NADP+ to proton translocation across a membrane. The enzyme comprises three components; dI binds NAD(H), dIII binds NADP(H), and dII spans the membrane. The 1,4,5,6-tetrahydro analogue of NADH (designated H2NADH) bound to isolated dI from Rhodospirillum rubrum transhydrogenase with similar affinity to the physiological nucleotide. Binding of either NADH or H2NADH led to closure of the dI mobile loop. The 1,4,5,6-tetrahydro analogue of NADPH (H2NADPH) bound very tightly to isolated R. rubrum dIII, but the rate constant for dissociation was greater than that for NADPH. The replacement of NADP+ on dIII either with H2NADPH or with NADPH caused a similar set of chemical shift alterations, signifying an equivalent conformational change. Despite similar binding properties to the natural nucleotides, neither H2NADH nor H2NADPH could serve as a hydride donor in transhydrogenation reactions. Mixtures of dI and dIII form dI2dIII1 complexes. The nucleotide charge distribution of complexes loaded either with H2NADH and NADP+ or with NAD+ and H2NADPH should more closely mimic the ground states for forward and reverse hydride transfer, respectively, than previously studied dead-end species. Crystal structures of such complexes at 2.6 and 2.3 A resolution are described. A transition state for hydride transfer between dihydronicotinamide and nicotinamide derivatives determined in ab initio quantum mechanical calculations resembles the organization of nucleotides in the transhydrogenase active site in the crystal structure. Molecular dynamics simulations of the enzyme indicate that the (dihydro)nicotinamide rings remain close to a ground state for hydride transfer throughout a 1.4 ns trajectory.  相似文献   

16.
Oxidation of NADPH catalyzed by the peroxidase/H2O2 system is known to require the presence of mediating molecules. Using either lactoperoxidase or horseradish peroxidase, we demonstrated that in the peroxidase/H2O2 system, NADPH oxidation was mediated by iodide. The oxidation product was the iodinated NADP. This product was shown to possess spectral characteristics different from those of NADP+ and NADPH, since for iodinated NADP, increased absorbance was observed in the 280-nm region and was directly proportional to the rate of iodination. It is suggested that oxidation and iodination of NADPH proceed via a single reaction between the intermediary iodide oxidation species and NADPH. Experiments with different molecules of NADPH analogues indicated that iodination occurred in the nicotinamide part of the NADPH molecule.  相似文献   

17.
Hypochlorous acid (HOCl) is formed by the action of the enzyme myeloperoxidase on hydrogen peroxide and chloride ions. It has been shown to be highly bactericidal and cytotoxic by a variety of mechanisms, one of which, may be the modification of DNA. Previously we have demonstrated by GC-MS analysis that exposure of calf thymus DNA to HOCl causes extensive pyrimidine modification, including 5-chlorocytosine formation. Using GC-MS analysis, we now demonstrate the formation of an additional chlorinated base product, 8-Cl adenine. The addition of 50 μM HOCl was sufficient to produce a significant increase in this product. The reaction of HOCl with adenine in calf thymus DNA was shown to be rapid with the reaction complete after 1 min. pH-dependence studies suggest HOCl rather than its conjugate base (OCl-) to be responsible for 8-Cl adenine formation. Other commercially available chlorinated base products, 6-Cl guanine or 2-Cl adenine were not detected. Therefore, 8-Cl adenine might prove a useful biomarker for studying the role of reactive chlorine species (RCS) during inflammatory processes.  相似文献   

18.
It has been found that melatonin reacts rapidly with hypochlorous acid in phosphate-buffered, ethanol-water solutions to produce 2-hydroxymelatonin. The rate law, d[2 - HOMel]/dt - kHOCl[Mel][HOCl] - kOCl-[Mel][OCl-], was obtained. At 37 degrees C and at a water concentration of 23.5 M, kOCl- = 6.0 x 10(2) L. mol-1. s-1, and kHOCl was found to be a function of the water concentration, kHOCl = 11 +/- 3 L3. mol-3. s-1. [H2O]2, indicating that the availability of water at the site of the reaction plays a significant role. The part that the structural components of melatonin play in determining the reaction pathway was examined by comparing the rate of deactivation of HOCl by melatonin to that of the model compounds indole, 5-methoxyindole, and 3-methylindole. The relative reactivity is explained in terms of steric and electronic effects, and it was found that the presence of the substituent at the 3-position influences the nature of the oxidation product. Melatonin and 3-methylindole yielded hydroxylated products, whereas indole and 5-methoxyindole produce chlorinated products.  相似文献   

19.
This study investigated the functional and structural effects of bovine Cu,Zn-superoxide dismutase (Cu,Zn-SOD) oxidation by the myeloperoxidase (MPO)/hydrogen peroxide (H 2 O 2 )/chloride system and reagent hypochlorous acid (HOCl). Exposure to HOCl led to a fast inactivation accompanied by structural alterations. The residual SOD activity depended on the reactants concentration ratio and on the exposure time. The concomitant high consumption of HOCl indicated the presence of multiple targets on the protein. As assessed by SDS/PAGE, HOCl caused the dissociation of the protein into protomers at 16 kDa stable to both SDS and reducing conditions. Results from isoelectric focusing gels showed that exposure to HOCl induced the formation of modified protein derivatives, with a more acidic net electric charge than the parent molecule, consistent with the presence of additional ions observed in the electrospray ionization mass spectra. The reaction of protein with HOCl resulted in changes in protein conformation as assessed by the UV fluorescence and oxidation of the unique methionine and tyrosine, chlorination of several lysines with formation of chloramines. There was no significant formation of dityrosine and carbonyl groups. Exposure to high levels of HOCl resulted in complete enzyme inactivation, loss of additional lysine, histidine and arginine residues and coincident detection of weakly bound zinc and copper using 4-pyridylazaresorcinol. Collectively, the results suggest that the decrease of the dismutase activity is probably related to both dissociation into protomers and unfolding due to extensive oxidative modifications of amino acids.  相似文献   

20.
This study investigated the functional and structural effects of bovine Cu,Zn-superoxide dismutase (Cu,Zn-SOD) oxidation by the myeloperoxidase (MPO)/hydrogen peroxide (H 2 O 2 )/chloride system and reagent hypochlorous acid (HOCl). Exposure to HOCl led to a fast inactivation accompanied by structural alterations. The residual SOD activity depended on the reactants concentration ratio and on the exposure time. The concomitant high consumption of HOCl indicated the presence of multiple targets on the protein. As assessed by SDS/PAGE, HOCl caused the dissociation of the protein into protomers at 16 kDa stable to both SDS and reducing conditions. Results from isoelectric focusing gels showed that exposure to HOCl induced the formation of modified protein derivatives, with a more acidic net electric charge than the parent molecule, consistent with the presence of additional ions observed in the electrospray ionization mass spectra. The reaction of protein with HOCl resulted in changes in protein conformation as assessed by the UV fluorescence and oxidation of the unique methionine and tyrosine, chlorination of several lysines with formation of chloramines. There was no significant formation of dityrosine and carbonyl groups. Exposure to high levels of HOCl resulted in complete enzyme inactivation, loss of additional lysine, histidine and arginine residues and coincident detection of weakly bound zinc and copper using 4-pyridylazaresorcinol. Collectively, the results suggest that the decrease of the dismutase activity is probably related to both dissociation into protomers and unfolding due to extensive oxidative modifications of amino acids.  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