首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The reaction of human myeloperoxidase (MPO) with hypochlorous acid (HOCl) was investigated by conventional stopped-flow spectroscopy at pH 5, 7, and 9. In the reaction of MPO with HOCl, compound I is formed. Its formation is strongly dependent on pH. HOCl (rather than OCl-) reacts with the unprotonated enzyme in its ferric state. Apparent second-order rate constants were determined to be 8.1 x 10(7) M(-1)s(-1) (pH 5), 2.0 x 10(8) M(-1)s(-1) (pH 7) and 2.0 x 10(6) M(-1)s(-1) (pH 9) at 15 degrees C. Furthermore, the kinetics and spectra of the reactions of halides and thiocyanate and of physiologically relevant one-electron donors (ascorbate, nitrite, tyrosine and hydrogen peroxide) with this compound I were investigated using the sequential-mixing technique. The results show conclusively that the redox intermediates formed upon addition of either hydrogen peroxide or hypochlorous acid to native MPO exhibit the same spectral features and reactivities and thus are identical. In stopped-flow investigations, the MPO/HOCl system has some advantage since: (i) in contrast to H2O2, HOCl cannot function as a one-electron donor of compound I; and (ii) MPO can easily be prevented from cycling by addition of methionine as HOCl scavenger. As a consequence, the observed absorbance changes are bigger and errors in data analysis are smaller.  相似文献   

2.
Abstract

The neutrophil enzyme myeloperoxidase catalyzes the oxidation of tyrosine to tyrosyl radicals, which cross-link to proteins and initiate lipid peroxidation. Tryptophan is present in plasma at about the same concentration as tyrosine and has a similar one-electron reduction potential. In this investigation, we have determined the ability of myeloperoxidase to catalyze the oxidation of tryptophan to assess whether or not this reaction may contribute to oxidative stress at sites of inflammation. We show that tryptophan is a poor substrate for myeloperoxidase because, even though it reacts rapidly with compound I (kI 2.1×106 M-1s-1), it reacts sluggishly with compound II (kII 7 M-1s-1). Tryptophan reversibly inhibited production of hypochlorous acid by purified myeloperoxidase by converting the enzyme to a mixture of compound II and compound III. It gave 50% inhibition (I50) at a concentration of 2 µM. In contrast, it was an ineffective inhibitor of hypochlorous acid production by human neutrophils (I50 80 µM) unless superoxide dismutase was present (I50 5 µM). We propose that compound I of myeloperoxidase will oxidize tryptophan at sites of inflammation. Enzyme turnover will result from the reaction of superoxide or tyrosine with compound II. Thus, tryptophan radicals are potential candidates for exacerbating oxidative stress during inflammation.  相似文献   

3.
Abstract

Myeloperoxidase is very susceptible to reducing radicals because the reduction potential of the ferric/ferrous redox couple is much higher compared with other peroxidases. Semiquinone radicals are known to reduce heme proteins. Therefore, the kinetics and spectra of the reactions of p-hydroquinone, 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone with compounds I and II were investi-gated using both sequential-mixing stopped-flow techniques and conventional spectrophotometric measurements. At pH 7 and 15°C the rate constants for compound I reacting with p-hydroquinone, 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone were determined to be 5.6±0.4×107 M-1s-1, 1.3±0.1×106 M-1s-1 and 3.1±0.3×106 M-1s-1, respectively. The corresponding reaction rates for compound II reduction were calculated to be 4.5±0.3×106 M-1s-1, 1.9±0.1×105 M-1s-1 and 4.5±0.2×104 M-1s-1, respectively. Semiquinone radicals, produced by compounds I and II in the classical peroxidation cycle, promote compound III (oxymyeloperoxidase) formation. We could monitor formation of ferrous myeloperoxidase as well as its direct transition to compound III by addition of molecular oxygen. Formation of ferrous myeloperoxidase is shown to depend strongly on the reduction potential of the corresponding redox couple benzoquinone/semiquinone. With 2,3-dimethylhydroquinone and 2,3,5-trimethylhydroquinone as substrate, myeloperoxidase is extremely quickly trapped as compound III. These MPO-typical features could have potential in designing specific drugs which inhibit the production of hypochlorous acid and consequently attenuate inflammatory tissue damage.  相似文献   

4.
Apocynin has been reported to require dimerization by myeloperoxidase (MPO) to inhibit leukocyte NADPH oxidase. (-)-Epicatechin, a dietary flavan-3-ol, has been identified as a ‘prodrug’ of apocynin-like metabolites that inhibit endothelial NADPH oxidase activity and elevate the cellular level of nitric oxide. Since (-)-epicatechin has tentatively been identified as substrate of MPO, we studied the one-electron oxidation of (-)-epicatechin by MPO. By using multi-mixing stopped-flow technique, we demonstrate that (-)-epicatechin is one of the most efficient electron donors for heme peroxidases investigated so far. Second order rate constants for the (-)-epicatechin-mediated conversion of MPO-compound I to compound II and compound II to resting enzyme were estimated to be 1.9 × 107 and 4.5 × 106 M−1 s−1, respectively (pH 7, 25 °C). The data indicate that (-)-epicatechin is capable of undergoing fast MPO-mediated one-electron oxidation.  相似文献   

5.
《Free radical research》2013,47(4):241-253
We have evaluated the abilities of ferulic acid, (±) catechin, (+) catechin and (-) epicatechin to scavenge the reactive oxygen species hydroxyl radical (OH±), hypochlorous acid (HOCl) and peroxyl radicals (RO2).

Ferulic acid tested at concentrations up to 5 mM inhibited the peroxidation of phospholipid liposomes. Both (±) and (+) catechin and (-) epicatechin were much more effective. All the compounds tested reacted with trichloromethyl peroxyl radical (CCl3O2) with rate constants > 1 × 106M?1s?1.

A mixture of FeCl3-EDTA, hydrogen peroxide (H2O2) and ascorbic acid at pH 7.4, has often been used to generate hydroxyl radicals (OH.) which are detected by their ability to cause damage to the sugar deoxyribose. Ferulic acid, (+) and (±) catechin and (-) epicatechin inhibited deoxyribose damage by reacting with OH. with rate constants of 4.5 × 109M?1s?1, 3.65 × 109M?1s?1, 2.36 × 109M?1s?1 and 2.84 × 109M?1s?1 respectively. (-) Epicatechin, ferulic acid and the (+) and (±) catechins exerted pro-oxidant action, accelerating damage to DNA in the presence of a bleomycin-iron complex. On a molar basis, ferulic acid was less effective in causing damage to DNA compared with the catechins.

A mixture of hypoxanthine and xanthine oxidase generates O2 which reduces cytochrome c to ferrocytochrome c. (+) Catechin and (-) epicatechin inhibited the reduction of cytochrome c in a concentration dependent manner. Ferulic acid and (±) catechin had only weak effects.

All the compounds tested were able to scavenge hypochlorous acid at a rate sufficient to protect alpha-1-antiproteinase against inactivation. Our results show that catechins and ferulic acid possess antioxidant properties. This may become important given the current search for “natural” replacements for synthetic antioxidant food additives.  相似文献   

6.
A key function of neutrophil myeloperoxidase (MPO) is the synthesis of hypochlorous acid (HOCl), a potent oxidizing agent that plays a cytotoxic role against invading bacteria and viruses at inflammatory sites and in phagosomes. MPO displayed a chlorinating activity preferably at acidic pH but at neutral pH MPO catalyzes mainly reactions of the peroxidase cycle. In the present work effects of tyrosine on the chlorinating activity of MPO were studied. At pH 7.4 we detected an increased HOCl production in the presence of tyrosine not only by the MPO-H2O2-Cl- system but also in suspensions of zymosan-activated neutrophils. An excess of H2O2 is known to cause an accumulation of compound II of MPO blocking the generation of HOCl at neutral pH. As evidenced by spectral changes, tyrosine-induced activation of MPO to synthesize HOCl was due to the ability of tyrosine to reduce compound II back to the native state, thus accelerating the enzyme turnover. MPO-induced oxidation of tyrosine is relevant to what can be in vivo; we detected MPO-catalyzed formation of dityrosine in the presence of plasma under experimental conditions when tyrosine concentration was about three magnitudes of order less than the Cl concentration. At acidic pH formation of compound II was impaired in the presence of chloride and dityrosine couldn't be detected in plasma. In conclusion, the ability of tyrosine to increase the chlorinating activity of MPO at neutral pH and enhanced values of H2O2 may be very effective for the specific enhancement of HOCl production under acute inflammation.  相似文献   

7.
Activated white cells use oxidants generated by the heme enzyme myeloperoxidase to kill invading pathogens. This enzyme utilizes H2O2 and Cl, Br, or SCN to generate the oxidants HOCl, HOBr, and HOSCN, respectively. Whereas controlled production of these species is vital in maintaining good health, their uncontrolled or inappropriate formation (as occurs at sites of inflammation) can cause host tissue damage that has been associated with multiple inflammatory pathologies including cardiovascular diseases and cancer. Previous studies have reported that sulfur-containing species are major targets for HOCl but as the reactions are fast the only physiologically relevant kinetic data available have been extrapolated from data measured at high pH (>10). In this study these values have been determined at pH 7.4 using a newly developed competition kinetic approach that employs a fluorescently tagged methionine derivative as the competitive substrate (k(HOCl + Fmoc-Met), 1.5×108 M−1 s−1). This assay was validated using the known k(HOCl + NADH) value and has allowed revised k values for the reactions of HOCl with Cys, N-acetylcysteine, and glutathione to be determined as 3.6×108, 2.9×107, and 1.24×108 M−1 s−1, respectively. Similar experiments with methionine derivatives yielded k values of 3.4×107 M−1 s−1 for Met and 1.7×108 M−1 s−1 for N-acetylmethionine. The k values determined here for the reaction of HOCl with thiols are up to 10-fold higher than those previously determined and further emphasize the critical importance of reactions of HOCl with thiol targets in biological systems.  相似文献   

8.
The reaction between ligninase and hydrogen peroxide yielding Compound I has been investigated using a stopped-flow rapid-scan spectrophotometer. The optical absorption spectrum of Compound I appears different to that reported by Andrawis, A. et al. (1987) and Renganathan, V. and Gold, M.H. (1986), in that the Soret-maximum is at 401 nm rather than 408 nm. The second-order rate constant (4.2·105 M−1·s−1) for the formation of Compound I was independent of pH (pH 3.0–6.0). In the absence of external electron donors, Compound I decayed to Compound II with a half-life of 5–10 s at pH 3.1. The rate of this reaction was not affected by the H2O2 concentration used. In the presence of either veratryl alcohol or ferrocyanide, Compound II was rapidly generated. With ferrocyanide, the second-order rate constant increased from 1.9·104 M−1·s−1 to 6.8·106 M−1·s−1 when the pH was lowered from 6.0 to 3.1. With veratryl alcohol as an electron donor, the second-order rate constant for the formation of Compound II increased from 7.0·103 M−1·s−1 at pH 6.0 to 1.0·105 M−1·s−1 at pH 4.5. At lower pH values the rate of Compound II formation no longer followed an exponential relationship and the steady-state spectral properties differed to those recorded in the presence of ferrocyanide. Our data support a model of enzyme catalysis in which veratryl alcohol is oxidized in one-electron steps and strengthen the view that veratryl alcohol oxidation involves a substrate-modified Compound II intermediate which is rapidly reduced to the native enzyme.  相似文献   

9.
Recombination of apo horseradish peroxidase with 2,4 dimethyldeutero hemin and its mono- and dimethyl esters was performed. The number of free carboxyl side chains in these three hemins is 2, 1 and 0 respectively. Despite such a difference, all of these three reconstituted enzymes can react with H2O2 to produce compound I. The second order rate constants for compound I formation are 1.3 × 107 M?1s?1, 8.5 × 106 M?1s?1 and 5.9 × 106 M?1s?1. Therefore the propionate side chain of hemin has no direct role in compound I formation.  相似文献   

10.
The heme-containing enzyme myeloperoxidase (MPO) accumulates at inflammatory sites and is able to catalyse one- and two-electron oxidation reactions. Here it is shown that (-)-epicatechin, which is known to have numerous beneficial health effects, in low micromolar concentration enhances the degradation of monochlorodimedon (MCD) or the chlorination of taurine in a concentration-dependent bell-shaped manner whereas at higher concentrations it sufficiently suppresses the release of hypochlorous acid. Presented reaction mechanisms demonstrate the efficiency of micromolar concentrations of the flavan-3-ol in overcoming the accumulation of compound II that does not participate in the chlorination cycle. In case of MCD the mechanism is more complicated since it also acts as peroxidase substrate with very different reactivity towards compound I (3 × 105 M−1 s−1) and compound II (8.8 M−1 s−1) at pH 7. By affecting the chlorinating activity of myeloperoxidase (-)-epicatechin may participate in regulation of immune responses at inflammatory sites.  相似文献   

11.
Abstract

The first complete mechanistic analysis of halide ion oxidation by a peroxidase was that of iodide oxidation by horseradish peroxidase. It was shown conclusively that a two-electron oxidation of iodide by compound I was occurring. This implied that oxygen atom transfer was occurring from compound I to iodide, forming hypoiodous acid, HOI. Searches were conducted for other two-electron oxidations. It was found that sulfite was oxidized by a two-electron mechanism. Nitrite and sulfoxides were not. If a competing substrate reduces some compound I to compound II by the usual one-electron route, then compound II will compete for available halide. Thus compound II oxidizes iodide to an iodine atom, I·, although at a slower rate than oxidation of I- by compound I. An early hint that mammalian peroxidases were designed for halide ion oxidation was obtained in the reaction of lactoperoxidase compound II with iodide. The reaction was accelerated by excess iodide, indicating a co-operative effect. Among the heme peroxidases, only chloroperoxidase (for example from Caldariomyces fumago) and mammalian myeloperoxidase are able to oxidize chloride ion. There is not yet a consensus as to whether the chlorinating agent produced in a peroxidase-catalyzed reaction is hypochlorous acid (HOCl), enzyme-bound hypochlorous acid (either Fe–HOCl or X–HOCl where X is an amino acid residue), or molecular chlorine Cl2. A study of the non-enzymatic iodination of tyrosine showed that the iodinating reagent was either HOI or I2. It was impossible to tell which species because of the equilibria:

I2+H2O=HOI+I-+H+</ p>

I-+I2=I3-

The same considerations apply to product analysis of an enzyme-catalyzed reaction. Detection of molecular chlorine Cl2 does not prove it is the chlorinating species. If Cl2 is in equilibrium with HOCl then one cannot tell which (if either) is the chlorinating reagent. Examples will be shown of evidence that peroxidase-bound hypochlorous acid is the chlorinating agent. Also a recent clarification of the mechanism of reaction of myeloperoxidase with hydrogen peroxide and chloride along with accurate determination of the elementary rate constants will be discussed.  相似文献   

12.
Abstract

Calf thymus DNA was treated with authentic HOCl, and hypohalous acid-generating systems. This caused a decrease in fluorescence of ethidium-DNA complexes when ethidium bromide was subsequently added to the DNA. The fluorescence continued to decrease up to 30 min after adding HOCl. Loss in fluorescence was proportional to the concentration of HOCl and was complete when a 3-fold excess of HOCl was added to the DNA. No significant decrease in the fluorescence was observed when the chlorination was carried out in the presence of a concentration of monochlorodimedone (MCD) equivalent to that of HOCl. MCD is known to react stoichiometrically with HOCl. The decrease in fluorescence was completely inhibited by H2O2, ascorbate and glutathione (GSH). We have estimated the rate constant for the reaction of HOCl with H2O2 to be 1–2×105 M-1s-1. When compared with authentic HOCl, HOCl-generating systems (Cl-+H2O2+MPO or chloroperoxidase) were found to be inefficient in damaging DNA. This result most likely arises because the rate constant for reaction of HOCl with H2O2 is about 1000-fold faster than that for the reaction with DNA. HOBr and HOI generating systems also had a limited ability to damage DNA. We conclude that good chlorine acceptors and antioxidants protect DNA from hypohalous acid-induced oxidative damage.  相似文献   

13.
Hypochlorous acid (HOCl) and N-chloramines are produced by myeloperoxidase (MPO) as part of the immune response to destroy invading pathogens. However, MPO also plays a detrimental role in inflammatory pathologies, including atherosclerosis, as inappropriate production of oxidants, including HOCl and N-chloramines, causes damage to host tissue. Low molecular mass thiol compounds, including glutathione (GSH) and methionine (Met), have demonstrated efficacy in scavenging MPO-derived oxidants, which prevents oxidative damage in vitro and ex vivo. Selenium species typically have greater reactivity toward oxidants compared to the analogous sulfur compounds, and are known to be efficient scavengers of HOCl and other hypohalous acids produced by MPO. In this study, we examined the efficacy of a number of sulfur and selenium compounds to scavenge a range of biologically relevant N-chloramines and oxidants produced by both isolated MPO and activated neutrophils and characterized the resulting selenium-derived oxidation products in each case. A dose-dependent decrease in the concentration of each N-chloramine was observed on addition of the sulfur compounds (cysteine, methionine) and selenium compounds (selenomethionine, methylselenocysteine, 1,4-anhydro-4-seleno-L-talitol, 1,5-anhydro-5-selenogulitol) studied. In general, selenomethionine was the most reactive with N-chloramines (k2 0.8–3.4×103 M–1 s–1) with 1,5-anhydro-5-selenogulitol and 1,4-anhydro-4-seleno-L-talitol (k2 1.1–6.8×102 M–1 s–1) showing lower reactivity. This resulted in the formation of the respective selenoxides as the primary oxidation products. The selenium compounds demonstrated greater ability to remove protein N-chloramines compared to the analogous sulfur compounds. These reactions may have implications for preventing cellular damage in vivo, particularly under chronic inflammatory conditions.  相似文献   

14.
Experimental evidence is provided that selenomethionine oxide (MetSeO) is more readily reducible than its sulfur analogue, methionine sulfoxide (MetSO). Pulse radiolysis experiments reveal an efficient reaction of MetSeO with one-electron reductants, such as e-aq (k = 1.2 × 1010M-1s-1), CO·-2 (k = 5.9 × 108 M-1s-1) and (CH3)2) C·OH (k = 3.5 × 107M-1s-1), forming an intermediate selenium-nitrogen coupled zwitterionic radical with the positive charge at an intramolecularly formed Se N 2σ/1σ* three-electron bond, which is characterized by an optical absorption with λmax at 375 nm, and a half-life of about 70 μs. The same transient is generated upon HO· radical-induced one-electron oxidation of selenomethionine (MetSe). This radical thus constitutes the redox intermediate between the two oxidation states, MetSeO and MetSe. Time-resolved optical data further indicate sulfur-selenium interactions between the Se N transient and GSH. The Se N transient appears to play a key role in the reduction of selenomethionine oxide by glutathione.  相似文献   

15.
In this study, we show that boronates, a class of synthetic organic compounds, react rapidly and stoichiometrically with peroxynitrite (ONOO) to form stable hydroxy derivatives as major products. Using a stopped-flow kinetic technique, we measured the second-order rate constants for the reaction with ONOO, hypochlorous acid (HOCl), and hydrogen peroxide (H2O2) and found that ONOO reacts with 4-acetylphenylboronic acid nearly a million times (k = 1.6 × 106 M− 1 s− 1) faster than does H2O2 (k = 2.2 M− 1 s− 1) and over 200 times faster than does HOCl (k = 6.2 × 103 M− 1 s− 1). Nitric oxide and superoxide together, but not alone, oxidized boronates to the same phenolic products. Similar reaction profiles were obtained with other boronates. Results from this study may be helpful in developing a novel class of fluorescent probes for the detection and imaging of ONOO formed in cellular and cell-free systems.  相似文献   

16.
《FEBS letters》1986,200(1):226-230
Magnesium binding to cation-depleted blue bacteriorhodopsin (b-bR) was studied spectrophotometrically as well as by following stopped-flow kinetics. There exist three kinetically different steps in the binding process, yielding purple bacteriorhodopsin (p-bR). Since only the firtst step is dependent on the concentration of the reactants, the reaction scheme
can be proposed as the simplest model, with MgbR being the first intermediate and ΣI denoting a set of successive intermediates. According to this model k1, k−1 and k2 are calculated to be 2.8 × 104 M−1 · s−1, 5.0 × 10 s−1 and 1 × 10−2 s−1, respectively.  相似文献   

17.
Flavonoids, plant polyphenols, ubiquitous components of human diet, are excellent antioxidants. Hypochlorous acid (HOCl), produced by activated neutrophils, is highly reactive chlorinating and oxidizing species. It has been reported earlier that flavonoids are chlorinated by HOCl. Here we show that flavonoids from flavonol subclass are also oxidized by HOCl, but only if the latter is in a large molar excess (≥?10). The kinetics of this reaction was studied by stopped-flow spectrophotometry, at different pH. We found that flavonols were oxidized by HOCl with the rate constants of the order of 104–105 M?1 s?1 at pH 7.5. Antioxidant activity of HOCl-modified flavonoids was measured by 2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) method. Slightly higher antioxidant activity, compared to parent compounds, was observed for flavonols after their reaction with equimolar or moderate excess of HOCl whereas flavonols treated with high molar excess of HOCl exhibited decrease in antioxidant activity. The mechanism of flavonoid reaction with HOCl at physiological pH is proposed, and biological consequences of this reaction are discussed.  相似文献   

18.
The kinetics of CNProto- and CNDeutero-hemin binding to apohemoglobin A2 was investigated in a stopped-flow device in 0.05 M potassium phosphate buffer, pH 7, at 10°C. The overall kinetic profile exhibited multiple phases: Phases I–IV corresponding with heme insertion (8.5?13 × 107 M?1 s?1), local structural rearrangement (0.21?0.23 s?1), global αδ structural event (0.071?0.098 s?1), and formation of the Fe–His bond (0.009?0.012 s?1), respectively. Kinetic differences observed between apohemoglobin A2 and apohemoglobin A (previously studied) prompted an analysis of the structures of β and δ chains through molecular modeling. This revealed a structural repositioning of the residues not only at, but also distant from the site of the amino acid substitutions, specifically those involved in the heme contact and subunit interface. A significant global change was observed in the structure of the exon-coded 3 region and provided additional evidence for the designation of this as the subunit assembly domain.  相似文献   

19.
Laccase-catalyzed oxidation of N-substituted phenothiazines and N-substituted phenoxazines was investigated at pH 5.5 and 25°C. The recombinant laccase from Polyporus pinsitus (rPpL) and the laccase from Myceliophthora thermophila (rMtL) were used. The dependence of initial reaction rate on substrate concentration was analyzed by applying the laccase action scheme in which the laccase native intermediate (NI) reacts with a substrate forming reduced enzyme. The reduced laccase produces peroxide intermediate (PI) which in turn decays to the NI. The calculated constant (kox) values of the PI formation are (6.1±3.1)×105 M−1s−1 for rPpL and (2.5±0.9)×104 M−1s−1 for rMtL. The bimolecular constants of the reaction of the native intermediate with electron donor (kred) vary in the interval from 2.2×105 to 2.1×107 M−1s−1 for rPpL and from 1.3×102 to 1.8×105 M-1s−1 for rMtL. The larger reactivity of rPpL in comparison to rMtL is associated with the higher redox potential of type I Cu of rPpL. The variation of kred values for both laccases correlates with the change of the redox potential of substrates. Following outer sphere (Marcus) electron transfer mechanism the calculated activationless electron transfer rate and the apparent reorganization energy are 5.0×107 M−1s−1 and 0.29 eV, respectively.  相似文献   

20.
D.B. Roberts 《FEBS letters》1983,156(1):193-196
Interaction of peanut agglutinin with MeUmbβGalβ(1→3)GalNAc was followed with the stopped-flow technique. The mechanism is a simple bimolecular association with k+ = 7.1 × 103 M?1. s?1 and k? = 0.24 s?1 at 25°C. The very slow dissociation rate of the complex strongly supports earlier conclusions that the combining site of peanut agglutinin is complementary to the Galβ(1→3)GalNac structure.  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