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.
Incubation of the nonphosphorylated form of maize-leaf phosphoenolpyruvate carboxylase (orthophosphate: oxaloacetate carboxy-lyase (phosphorylating), PEPC, EC 4.1.1.31) with the reagent pyridoxal 5′-phosphate (PLP) resulted in time-dependent, reversible inactivation and desensitization to the activator glucose 6-phosphate (Glc6P) and other related phosphorylated compounds. Both processes are not connected, since (i) when the PLP-modification was carried out in the presence of saturating ligands of the active site, which prevents inactivation, the desensitization to Glc6P is still observed, and (ii) under some experimental conditions the desensitization reaction is 4-times faster than the inactivation. Desensitization to Glc6P is first order with respect to PLP and has a second-order forward rate constant of 4.7±0.3 s−1 M−1 and a first-order reverse rate constant of 0.0046±0.0002 s−1. Correlation studies between the remaining Glc6P sensitivity and mol of PLP residues incorporated per mol of enzyme subunit indicate that one lysyl group for enzyme monomer is involved in the sensitivity of the enzyme to Glc6P. The reactivity of this group is increased by polyethylene glycol and glycerol, while the reactivity of the lysyl group of the active site is not affected by these organic cosolutes. In the presence but not in the absence of the organic cosolutes, Glc6P by itself offers significant protection against desensitization, while increases the extent of inactivation. Free PEP or PEP-Mg have opposite effects, protecting the enzyme against inactivation and increasing the degree of desensitization. They also increases the protection against desensitization afforded by Glc6P. Finally, the PEPC inhibitor malate provides some protection against both inactivation and desensitization. Taken together, these results are consistent with PLP-modification of a highly reactive lysyl group at or near the allosteric Glc6P-site.  相似文献   

2.
Tryptophan hydroxylase, the initial and rate-limiting enzyme in serotonin biosynthesis, is inactivated by peroxynitrite in a concentration-dependent manner. This effect is prevented by molecules that react directly with peroxynitrite such as dithiothreitol, cysteine, glutathione, methionine, tryptophan, and uric acid but not by scavengers of superoxide (superoxide dismutase), hydroxyl radical (Me(2)SO, mannitol), and hydrogen peroxide (catalase). Assuming simple competition kinetics between peroxynitrite scavengers and the enzyme, a second-order rate constant of 3.4 x 10(4) M(-1) s(-1) at 25 degrees C and pH 7.4 was estimated. The peroxynitrite-induced loss of enzyme activity was accompanied by a concentration-dependent oxidation of protein sulfhydryl groups. Peroxynitrite-modified tryptophan hydroxylase was resistant to reduction by arsenite, borohydride, and dithiothreitol, suggesting that sulfhydryls were oxidized beyond sulfenic acid. Peroxynitrite also caused the nitration of tyrosyl residues in tryptophan hydroxylase, with a maximal modification of 3.8 tyrosines/monomer. Sodium bicarbonate protected tryptophan hydroxylase from peroxynitrite-induced inactivation and lessened the extent of sulfhydryl oxidation while causing a 2-fold increase in tyrosine nitration. Tetranitromethane, which oxidizes sulfhydryls at pH 6 or 8, but which nitrates tyrosyl residues at pH 8 only, inhibited tryptophan hydroxylase equally at either pH. Acetylation of tyrosyl residues with N-acetylimidazole did not alter tryptophan hydroxylase activity. These data suggest that peroxynitrite inactivates tryptophan hydroxylase via sulfhydryl oxidation. Modification of tyrosyl residues by peroxynitrite plays a relatively minor role in the inhibition of tryptophan hydroxylase catalytic activity.  相似文献   

3.
Human recombinant copper-zinc superoxide dismutase (CuZnSOD) was inactivated by peroxynitrite, the product of the reaction between nitric oxide and superoxide. The concentration of peroxynitrite that decreased the activity by 50% (IC(50)) was approximately 100 microM at 5 microM CuZnSOD and the inactivation was higher at alkaline pH. Stopped-flow determinations showed that the second-order rate constant for the direct reaction of peroxynitrite with CuZnSOD was (9.4 +/- 1.0) x 10(3) M(-1) s(-1) per monomer at pH 7.5 and 37 degrees C. Addition of peroxynitrite (1 mM) to CuZnSOD (0.5 mM) in the presence of the spin trap 2-methyl-2-nitrosopropane led to the electron paramagnetic resonance detection of an anisotropic signal typical of a protein radical adduct. Treatment with Pronase revealed a nearly isotropic signal consistent with the formation of histidinyl radical. The effects of nitrite, hydrogen peroxide, bicarbonate, and mannitol on the inactivation were assessed. Considering the mechanism accepted for the reaction of CuZnSOD with hydrogen peroxide and the fact that CuZnSOD promotes the nitration of phenolics by peroxynitrite, we herein propose that peroxynitrite reacts with CuZnSOD leading to nitrogen dioxide plus a copper-bound hydroxyl radical species that reacts with histidine residues, forming histidinyl radical.  相似文献   

4.
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   

5.
Seleno-organic compounds are known as efficient “scavengers” of peroxynitrite (PN). Here we studied the protective effect of selenolipoic acid (SeLA), the seleno-containing analogue of lipoic acid, on peroxynitrite-dependent inactivation of NADPH-cytochrome P450 reductase. 3-Morpholinosydnonimine hydrochloride (SIN-1) was used as a source of peroxynitrite. The reductase was irreversibly inactivated by PN generated from SIN-1. The inactivation occurred with the rate constant of about 3 × 104M-1s-1. The presence of SeLA at low concentration (0.5 μM) led to synergistic increase of the reductase inactivation by PN. Our results suggest the formation of a reactive derivative of SeLA in the reaction of SeLA with PN, probably selenolseleninate, that mediates the aggravation of reductase inactivation. In the presence of SeLA, the inactivation was reversible under the action of thiols, allowing us to conclude that the observed action of SeLA may be considered as protective.  相似文献   

6.
Malic enzyme (S)-malate: NADP+ oxidoreductase (oxaloacetate-decar☐ylating, EC 1.1.1.40) purified from the thermoacidophilic archaebacterium Sulfolobus solfataricus, strain MT-4, catalyzed the metal-dependent decar☐ylation of oxaloacetate at optimum pH 7.6 at a rate comparable to the decar☐ylation of l-malate. The oxaloacetate decar☐ylase activity was stimulated about 50% by NADP but only in the presence of MgCl2, and was strongly inhibited by l-malate and NADPH which abolished the NADP activation. In the presence of MnCl2 and in the absence of NADP, the Michaelis constant and Vm for oxaloacetate were 1.7 mM and 2.3 μmol·min−1·mg−1, respectively. When MgCl2 replaced MnCl2, the kinetic parameters for oxaloacetate remained substantially unvaried, whereas the Km and Vm values for l-malate have been found to vary depending on the metal ion. The enzyme carried out the reverse reaction (malate synthesis) at about 70% of the forward reaction, at pH 7.2 and in the presence of relatively high concentrations of bicarbonate and pyruvate. Sulfhydryl residues (three cysteine residues per subunit) have been shown to be essential for the enzymatic activity of the Sulfolobus solfataricus malic enzyme. 5,5′-Dithiobis(2-nitrobenzoic acid), p-hydroxymercuribenzoate and N-ethylmaleimide caused the inactivation of the oxidative decar☐ylase activity, but at different rates. The inactivation of the overall activity by p-hydroxymercuribenzoate was partially prevented by NADP singly or in combination with both l-malate and MnCl2, and strongly enhanced by the car☐ylic acid substrates; NADP + malate + MnCl2 afforded total protection. The inactivation of the oxaloacetate decar☐ylase activity by p-hydroxymercuribenzoate treatment was found to occur at a slower rate than that of the oxidative decar☐ylase activity.  相似文献   

7.
Human copper-zinc superoxide dismutase undergoes inactivation when exposed to O2? and H2O2 generated during the oxidation of acetaldehyde by xanthine oxidase at pH 7.4 and 37° C. In contrast, human manganese superoxide dismutase is not inactivated under the same conditions. Catalase and Mn-superoxide dismutase protect CuZn superoxide dismutase from inactivation. Similar protection is observed with hydroxyl radical (OH.) scavengers, such as formate and mannitol. In contrast, other OH. scavengers such as ethanol and tert-butyl alcohol, have no protective action. The latter results indicate that “free OH.” is not responsible for the inactivation. Furthermore, H2O2 generated during the oxidation of glucose by glucose oxidase, i.e., without production of O2?, does not induce CuZn superoxide dismutase inactivation. A mechanism accounting for this O2?H2O2-dependent inactivation of CuZn superoxide dismutase is proposed.  相似文献   

8.
Chicken liver mevalonate 5-diphosphate decar☐ylase (ATP:(R)-5-diphosphomevalonate car☐y-lyase (dehydrating), EC 4.1.1.33) is inactivated by methylmethanethiosulfonate and 5,5′-dithiobis(2-nitrobenzoate). The presence of the substrates ATP or mevalonate 5-diphosphate protect very effectively against inactivation. The inactivation is second order with respect to methylmethanethiosulfonate, with an inactivation rate constant of (7.6 ± 0.1) · 10−5 μM−2 ·s−1, implying that the modifier may be reacting with more than one thiol in the enzyme. The enzyme is also inactivated by a number of dithiol-specific reagents, suggesting the presence of a functional dithiol. The determined pKapp values for enzyme modification by methyl methanethiosulfonate and phenylarsine oxide are 7.3 ± 0.1 and 7.6 ± 0.3, respectively. From the data presented, it is concluded that the enzyme posseses a functional dithiol that is important for substrate binding.  相似文献   

9.
Trihydroxybenzenes are degraded anaerobically through the phloroglucinol pathway. In Pelobacter acidigallici as well as in Pelobacter massiliensis, pyrogallol is converted to phloroglucinol in the presence of 1,2,3,5-tetrahydroxybenzene by intermolecular hydroxyl transfer. The enzyme catalyzing this reaction was purified to chromatographic and electrophoretic homogeneity. Gel filtration and electrophoresis revealed a heterodimer structure with an apparent molecular mass of 127 kDa for the native enzyme and 86 kDa and 38 kDa, respectively, for the subunits. The enzyme was not sensitive to oxygen. HgCl2, p-chloromercuribenzoic acid, and CuCl2 inhibited strongly the reaction indicating an essential function of SH-groups. Transhydroxylase had a pH-optimum of 7.0 and a pI of 4.1. The apparent temperature optimum was in the range of 53°C to 58°C. The activation energy for the conversion of pyrogallol and 1,2,3,5-tetrahydroxybenzene to phloroglucinol and tetrahydroxybenzene was 31.4 kJ per mol. Purified enzyme exhibited a specific activity of 3.1 mol. m−1 mg−1 protein and an apparent Km for pyrogallol and 1,2,3,5-tetrahydroxybenzene of 0.70 mM and 0.71 mM, respectively. The enzyme was found to contain per mol heterodimer 1.1 mol molybdenum, 12.1 mol iron and 14.5 mol acid-labile sulfur. Requirement for molybdenum for transhydroxylating enzyme activity was proven also by cultivation experiments. No hints for the presence of flavins were obtained. The results presented here support the hypothesis that a redox reaction is involved in this intermolecular hydroxyl transfer.  相似文献   

10.
Macrophage-derived nitric oxide (NO) participates in cytotoxic mechanisms against diverse microorganisms and tumor cells. These effects can be mediated by NO itself or NO-derived species such as peroxynitrite formed by its diffusion-controlled reaction with NADPH oxidase-derived superoxide radical anion (O2). In vivo, the facile extracellular diffusion of NO as well as different competing consumption routes limit its bioavailability for the reaction with O2 and, hence, peroxynitrite formation. In this work, we evaluated the extent by which NO diffusion to red blood cells (RBC) can compete with activated macrophages-derived O2 and affect peroxynitrite formation yields. Macrophage-dependent peroxynitrite production was determined by boron-based probes that react directly with peroxynitrite, namely, coumarin-7-boronic acid (CBA) and fluorescein-boronate (Fl-B). The influence of NO diffusion to RBC on peroxynitrite formation was experimentally analyzed in co-incubations of NO and O2-forming macrophages with erythrocytes. Additionally, we evaluated the permeation of NO to RBC by measuring the intracellular oxidation of oxyhemoglobin to methemoglobin. Our results indicate that diluted RBC suspensions dose-dependently inhibit peroxynitrite formation, outcompeting the O2 reaction. Computer-assisted kinetic studies evaluating peroxynitrite formation by its precursor radicals in the presence of RBC are in accordance with experimental results. Moreover, the presence of erythrocytes in the proximity of NO and O2--forming macrophages prevented intracellular Fl-B oxidation pre-loaded in L1210 cells co-cultured with activated macrophages. On the other hand, Fl-B-coated latex beads incorporated in the macrophage phagocytic vacuole indicated that intraphagosomal probe oxidation by peroxynitrite was not affected by nearby RBC. Our data support that in the proximity of a blood vessel, NO consumption by RBC will limit the extracellular formation (and subsequent cytotoxic effects) of peroxynitrite by activated macrophages, while the intraphagosomal yield of peroxynitrite will remain unaffected.  相似文献   

11.
The reactions of NO2 with both oxidized and reduced cytochrome c at pH 7.2 and 7.4, respectively, and with N-acetyltyrosine amide and N-acetyltryptophan amide at pH 7.3 were studied by pulse radiolysis at 23 °C. NO2 oxidizes N-acetyltyrosine amide and N-acetyltryptophan amide with rate constants of (3.1±0.3)×105 and (1.1±0.1)×106 M−1 s−1, respectively. With iron(III)cytochrome c, the reaction involves only its amino acids, because no changes in the visible spectrum of cytochrome c are observed. The second-order rate constant is (5.8±0.7)×106 M−1 s−1 at pH 7.2. NO2 oxidizes iron(II)cytochrome c with a second-order rate constant of (6.6±0.5)×107 M−1 s−1 at pH 7.4; formation of iron(III)cytochrome c is quantitative. Based on these rate constants, we propose that the reaction with iron(II)cytochrome c proceeds via a mechanism in which 90% of NO2 oxidizes the iron center directly—most probably via reaction at the solvent-accessible heme edge—whereas 10% oxidizes the amino acid residues to the corresponding radicals, which, in turn, oxidize iron(II). Iron(II)cytochrome c is also oxidized by peroxynitrite in the presence of CO2 to iron(III)cytochrome c, with a yield of ~60% relative to peroxynitrite. Our results indicate that, in vivo, NO2 will attack preferentially the reduced form of cytochrome c; protein damage is expected to be marginal, the consequence of formation of amino acid radicals on iron(III)cytochrome c.  相似文献   

12.
Peroxynitrite anion is a reactive and short-lived species and its formation in vivo has been implicated in several human diseases. In view of the potential usefulness of compounds that can protect against peroxynitrite or their reactive intermediates, a study focused on flavonoid compounds was carried out. Since the reactivity of peroxynitrite may be modified by [Formula: See Text] which is an important plasma buffer, the protection of flavonoids against peroxynitrite was evaluated by their ability to inhibit the peroxynitrite-mediated dihydrorhodamine (DHR123) oxidation with or without physiological concentrations of bicarbonate. Flavonoids from different classes were studied to elucidate which structural features are required for an effective protection. The most efficient flavonoids on protecting DHR123 against oxidation by peroxynitrite have their ability diminished in the presence of bicarbonate, but they maintain the hierarchy established in the absence of bicarbonate. The flavones are the most effective flavonoids and their effects depend mainly on the number of hydroxyl groups. These must include either a catechol group in the B-ring or a hydroxyl group at the 3-position. This work also included some isoflavones, flavanones and a flavanol, which enable us to conclude about the importance of another structural feature: the 2,3-double bond. These results indicate that the ability of flavonoids to protect against peroxynitrite depends on some structural features, also important to scavenge oxygen free radicals and to chelate metal ions. The most efficient flavonoids are effective at low concentrations with IC50 of the same magnitude as Ebselen, a selenocompound that has been reported to be excellent at protecting against peroxynitrite. Their effectiveness at low concentrations is an important aspect to take into account when characterizing a compound as an antioxidant with biological interest.  相似文献   

13.
R.R. Sharp  C.F. Yocum 《BBA》1981,635(1):90-104
The kinetics of Mn release during NH2OH inactivation of the water oxidizing reaction is largely insensitive to the S-state present during addition of NH2OH. This appears to reflect reduction by NH2OH of higher S-states to a common more reduced state (S0 or S?1) which alone is susceptible to NH2OH inactivation. Sequences of saturating flashes with dark intervals in the range 0.2–5 s?1 effectively prevent NH2OH inactivation and the associated liberation of manganese. This light-induced protection disappears rapidly when the dark interval is longer than about 5 s. Under continuous illumination, protection against NH2OH inactivation is maximally effective at intensities in the range 103–104 erg · cm?2 · s?1. This behavior differs from that of NH2OH-induced Mn release, which is strongly inhibited at all intensities greater than 103 erg · cm?2 · s?1. This indicates that two distinct processes are responsible for inactivation of water oxidation at high and low intensities. Higher S-states appear to be immune to the reaction by which NH2OH liberates manganese, although the overall process of water oxidation is inactivated by NH2OH in the presence of intense light. The light-induced protection phenomenon is abolished by 50 μM DCMU, but not by high concentrations of carbonyl cyanide m-chlorophenylhydrazone, which accelerates inactivation reactions of the water-splitting enzyme, Y (an ADRY reagent). The latter compound accelerates both inactivation of water oxidation and manganese extraction in the dark.  相似文献   

14.
Rabbit muscle pyruvate kinase is inactivated by 2,3-butanedione in borate buffer. The inactivation follows pseudo-first-order kinetics with a calculated second-order rate constant of 4.6 m?1 min?1. The modification can be reversed with almost total recovery of activity by elimination of the butanedione and borate buffer, suggesting that only arginyl groups are modified; this result agrees with the loss of arginine detected by amino acid analysis of the modified enzyme. Using the kinetic data, it was estimated that the reaction of a single butanedione molecule per subunit of the enzyme is enough to completely inactivate the protein. The inactivation is partially prevented by phosphoenolpyruvate in the presence of K+ and Mg2+, but not by the competitive inhibitors lactate and bicarbonate. These findings point to an essential arginyl residue being located near the phosphate binding site of phosphoenolpyruvate.  相似文献   

15.
The α-carbonic anhydrase gene from Helicobacter pylori strain 26695 has been cloned and sequenced. The full-length protein appears to be toxic to Escherichia coli, so we prepared a modified form of the gene lacking a part that presumably encodes a cleavable signal peptide. This truncated gene could be expressed in E. coli yielding an active enzyme comprising 229 amino acid residues. The amino acid sequence shows 36% identity with that of the enzyme from Neisseria gonorrhoeae and 28% with that of human carbonic anhydrase II. The H. pylori enzyme was purified by sulfonamide affinity chromatography and its circular dichroism spectrum and denaturation profile in guanidine hydrochloride have been measured. Kinetic parameters for CO2 hydration catalyzed by the H. pylori enzyme at pH 8.9 and 25°C are kcat=2.4×105 s−1, KM=17 mM and kcat/KM=1.4×107 M−1 s−1. The pH dependence of kcat/KM fits with a simple titration curve with pKa=7.5. Thiocyanate yields an uncompetitive inhibition pattern at pH 9 indicating that the maximal rate of CO2 hydration is limited by proton transfer between a zinc-bound water molecule and the reaction medium in analogy to other forms of the enzyme. The 4-nitrophenyl acetate hydrolase activity of the H. pylori enzyme is quite low with an apparent catalytic second-order rate constant, kenz, of 24 M−1 s−1 at pH 8.8 and 25°C. However, with 2-nitrophenyl acetate as substrate a kenz value of 665 M−1 s−1 was obtained under similar conditions.  相似文献   

16.
Summary

The reaction between peroxynitrous acid (hydrogen oxoperoxonitrate) and L-tryptophan is 130 M?1s?1 at 25°C. The pH dependence of the second-order rate constant shows a maximum at pH 5.1. The enthalpy and entropy of activation at pH 7.1 are 10.6 ± 0.4 kcal.mol?1 and -16 ± 2 cal.mol?1K?1 respectively. High-performance liquid chromatography analysis revealed a number of reaction products, two of which were identified as 5- and 6- nitrotryptophan. Hydroxytryptophans were not observed, even at low peroxynitrite concentrations where most of the peroxynitrite decays to nitrate via a first-order process. These results support the hypothesis that isomerization of protonated peroxynitrite to nitrate does not involve formation of the hydroxyl radical.  相似文献   

17.
The anation reaction of aquopentaamminecobalt(III) by acetate has been studied in the temperature range 60–80°C and acidity range 1.0 ≦ pH ≦ 5.5 for total acetate concentrations ≦ 0.5 M and at ionic strength 1.0 M. The anation by acetic acid follows second-order kinetics (k0), whereas the kinetic results for the anation by acetate (Q1, k1) provide evidence for the formation of an ion-pair with the complex ion. Typical experimental results at 70°C are k0 = 5.33 X 10−5 M−1 sec−1, Q1 = 5.87 M−1 and k1 = 1.46 X 10−4sec−1. The activation parameters for the different reaction paths are reported and the results discussed with reference to various other anation reactions of Co(III) complexes.  相似文献   

18.
The basic properties of purified d-amino acid oxidase from the yeast Trigonopsis variabilis were investigated. The pH optimum of activity was between pH 8.5 and 9.0, and the native molecular masses of holo- and apo-enzyme were determined to be 170 kDa; higher aggregates corresponded to molecular masses of 320 and 570 kDa. The apparent V max and K m values for different substrates varied between 3.7 to 185 U/mg and 0.2 to 17.3 mM, respectively. The reaction of d-amino acid oxidase with sulfite was followed by the typical spectral modifications of the FAD resembling the reduced enzyme; a K d of 30 μM was calculated for the N(5)-adduct. The red anionic flavin radical of the enzyme was stable; benzoate had no influence on the spectral properties. A complete loss of enzyme activity was observed after chemical modification by the histidine-specific reagent diethyl pyrocarbonate. The inactivation showed pseudo-first-order kinetics, with a second-order rate constant of 13.6 M–1 min–1 at pH 6.0 and 20°C. The addition of a substrate under anoxic conditions led to a substantial protection from inactivation, which indicates a localization of the modified residues close to the active site. The pKa of the reacting group was determined to be 7.7, and the rate of inactivation reached a limiting value of 0.031 min–1. Received: 22 August 1995 / Accepted: 17 October 1995  相似文献   

19.
It has been demonstrated that the NO produced by nitric oxide synthase or by the reduction of nitrite by nitrate reductase plays an important role in plants’ defense against microbial pathogens. The detection of nitrosyl Lb in nodules strongly suggests that NO is also formed in functional nodules. Moreover, NO may react with superoxide (which has been shown to be produced in nodules by various processes), leading to the formation of peroxynitrite. We have determined the second-order rate constants of the reactions of soybean oxyleghemoglobin with nitrogen monoxide and peroxynitrite. At pH 7.3 and 20 °C, the values are on the order of 108 and 104 M−1 s−1, respectively. In the presence of physiological amounts of CO2 (1.2 mM), the second-order rate constant of the reaction of oxyleghemoglobin peroxynitrite is even larger (105 M−1 s−1). The results presented here clearly show that oxyleghemoglobin is able to scavenge any NO and peroxynitrite formed in functional nodules. This may help to stop NO triggering a plant defense reaction.  相似文献   

20.
Peroxynitrite is implicated in many diseases. Hence, there is considerable interest in potential therapeutic peroxynitrite scavengers. Diet-derived phenolics have been claimed to be powerful peroxynitrite scavengers. However, the reactivity of peroxynitrite can be significantly modified by bicarbonate and this has not been considered in evaluations of the scavenging activity of phenols. Bicarbonate (25 mM) significantly decreased the ability of several phenolic compounds (caffeic acid, o- and p-coumaric acid, gallic acid, ferulic acid) but not others (catechin and epicatechin) to inhibit peroxynitrite-mediated tyrosine nitration. Bicarbonate (25 mM) also decreased the ability of catechin, epicatechin, quercetin and ferulic acid but not chlorogenic acid, gallic acid, caffeic acid and o-coumaric acid to inhibit peroxynitrite-mediated alpha(1)-antiproteinase inactivation. These results show that physiological concentrations of bicarbonate substantially modify the ability of dietary phenolics to prevent peroxynitrite-mediated reactions. When assessing compounds for peroxynitrite scavenging, experiments should be conducted in the presence of bicarbonate to avoid misleading results.  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