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.
Peroxidase catalyzed reactions of iodide at low pH   总被引:1,自引:0,他引:1  
Lactoperoxidase (LP) and horseradish peroxidase (HRP) catalyze the rapid oxidation of iodide to iodine at pH 3.6. One mole of peroxide reacts with 2 moles of iodide, producing 1 mole of iodine. Neither enzyme catalyzes the further oxidation of iodine. The turnover numbers for LP and HRP are 1.4 × 105 and 2.2 × 104 I2 moles produced/min/enzyme mole, respectively.  相似文献   

2.
The physiological function of urate is poorly understood. It may act as a danger signal, an antioxidant, or a substrate for heme peroxidases. Whether it reacts sufficiently rapidly with lactoperoxidase (LPO) to act as a physiological substrate remains unknown. LPO is a mammalian peroxidase that plays a key role in the innate immune defense by oxidizing thiocyanate to the bactericidal and fungicidal agent hypothiocyanite. We now demonstrate that urate is a good substrate for bovine LPO. Urate was oxidized by LPO to produce the electrophilic intermediates dehydrourate and 5-hydroxyisourate, which decayed to allantoin. In the presence of superoxide, high yields of hydroperoxides were formed by LPO and urate. Using stopped-flow spectroscopy, we determined rate constants for the reaction of urate with compound I (k1 = 1.1 × 107 m−1 s−1) and compound II (k2 = 8.5 × 103 m−1 s−1). During urate oxidation, LPO was diverted from its peroxidase cycle because hydrogen peroxide reacted with compound II to give compound III. At physiologically relevant concentrations, urate competed effectively with thiocyanate, the main substrate of LPO for oxidation, and inhibited production of hypothiocyanite. Similarly, hypothiocyanite-dependent killing of Pseudomonas aeruginosa was inhibited by urate. Allantoin was present in human saliva and associated with the concentration of LPO. When hydrogen peroxide was added to saliva, oxidation of urate was dependent on its concentration and peroxidase activity. Our findings establish urate as a likely physiological substrate for LPO that will influence host defense and give rise to reactive electrophilic metabolites.  相似文献   

3.
Allantoin in the presence of calcium ions has been implicated as a potential toxic agent in Reye's syndrome. An investigation of possible alternative sources of allantoin in humans, which lack the enzyme uricase, has been initiated. Urate is a strong reducing agent which can reduce cytochrome c nonenzymatically, with the concomitant production of CO2 and H+. The stoichiometries measured for the various reactants and products were 1 urate:2 cytochrome c:1 H+:1 CO2. The initial reaction rate depended on the concentrations of both urate and cytochrome c, with reaction kinetics that were first order with respect to urate and second order with respect to cytochrome c. The participation of molecular oxygen in this reaction could not be detected. The pH and ionic strength optima for this reaction were determined to be 9.5–10.5 and 10−5m, respectively. Based on the results reported here, the following balanced equation can be written: urate−2 + 2 cytochrome c+3 + 2 H2O → allantoin + 2 cytochrome c+2 + H+ + HCO3. We propose that allantoin can be generated from the oxidation of urate by cytochrome c+3, and that this is a potential source of allantoin in human tissues.  相似文献   

4.
Ascorbate is catalytically oxidized by a coupled iron-ceruloplasmin system, the iron ions functioning as a red/ox cycling intermediate between ceruloplasmin and ascorbate. Serum albumin, an iron binding compound, was found to stimulate the ascorbate oxidation rate. It is proposed that ferrous ions react more rapidly with ceruloplasmin when they are bound to albumin. A K m value of 39 m was estimated for Fe2+-albumin. Citrate and urate inhibit the iron-ceruloplasmin-dependent ascorbate oxidation by chelating ferric ions. In the presence of albumin only citrate reduced the oxidation rate, the observation suggesting the following order of iron binding ability: citrate > albumin > urate. Physiological aspects of the results have been discussed.  相似文献   

5.
Peroxisomes and reactive oxygen species,a lasting challenge   总被引:1,自引:0,他引:1  
Oxidases generating and enzymes scavenging H2O2 predestine peroxisomes (PO) to a pivotal organelle in oxygen metabolism. Catalase, the classical marker enzyme of PO, exhibits both catalatic and peroxidatic activity. The latter is responsible for the staining with 3,3′-diamino-benzidine, which greatly facilitated the visualization of the organelle and promoted further studies on PO. d-Amino acid oxidase catalyzes with strict stereospecificity the oxidative deamination of d-amino acids. The oxidase is significantly more active in the kidney than in liver and more in periportal than pericentral rat hepatocytes. Peroxisomes in these tissues differ in their enzyme activity and protein concentration not only in adjacent cells but even within the same one. Moreover, the enzyme appears preferentially concentrated in the central region of the peroxisomal matrix compartment. Urate oxidase, a cuproprotein catalyzing the oxidation of urate to allantoin, is confined to the peroxisomal core, yet is lacking in human PO. Recent experiments revealed that cores in rat hepatocytes appear in close association with the peroxisomal membrane releasing H2O2 generated by urate oxidase to the surrounding cytoplasma. Xanthine oxidase is exclusively located to cores, oxidizes xanthine thereby generating H2O2 and O2 radicals. The latter are converted to O2 and H2O2 by CuZn superoxide dismutase, which has been shown recently to be a bona fide peroxisomal protein. Presented at the 50th Anniversary Symposium of the Society for Histochemistry, Interlaken, Switzerland, October 1-4, 2008.  相似文献   

6.
BackgroundExtracellular surface protein disulfide isomerase-A1 (PDI) is involved in platelet aggregation, thrombus formation and vascular remodeling. PDI performs redox exchange with client proteins and, hence, its oxidation by extracellular molecules might alter protein function and cell response. In this study, we investigated PDI oxidation by urate hydroperoxide, a newly-described oxidant that is generated through uric acid oxidation by peroxidases, with a putative role in vascular inflammation.MethodsAmino acids specificity and kinetics of PDI oxidation by urate hydroperoxide was evaluated by LC-MS/MS and by stopped-flow. Oxidation of cell surface PDI and other thiol-proteins from HUVECs was identified using impermeable alkylating reagents. Oxidation of intracellular GSH and GSSG was evaluated with specific LC-MS/MS techniques. Cell adherence, detachment and viability were assessed using crystal violet staining, cellular microscopy and LDH activity, respectively.ResultsUrate hydroperoxide specifically oxidized cysteine residues from catalytic sites of recombinant PDI with a rate constant of 6 × 103 M−1 s−1. Incubation of HUVECs with urate hydroperoxide led to oxidation of cell surface PDI and other unidentified cell surface thiol-proteins. Cell adherence to fibronectin coated plates was impaired by urate hydroperoxide, as well as by other oxidants, thiol alkylating agents and PDI inhibitors. Urate hydroperoxide did not affect cell viability but significantly decreased GSH/GSSG ratio.ConclusionsOur results demonstrated that urate hydroperoxide affects thiol-oxidation of PDI and other cell surface proteins, impairing cellular adherence.General significanceThese findings could contribute to a better understanding of the mechanism by which uric acid affects endothelial cell function and vascular homeostasis.  相似文献   

7.
Lactoperoxidase is an enzyme that contributes to the antimicrobial defense in secretory fluids and that has attracted interest as a potential biopreservative for foods and other perishable products. Its antimicrobial activity is based on the formation of hypothiocyanate (OSCN) from thiocyanate (SCN), using H2O2 as an oxidant. To gain insight into the antibacterial mode of action of the lactoperoxidase enzyme system, we generated random transposon insertion mutations in Escherichia coli MG1655 and screened the resultant mutants for an altered tolerance of bacteriostatic concentrations of this enzyme system. Out of the ca. 5,000 mutants screened, 4 showed significantly increased tolerance, and 2 of these had an insertion, one in the waaQ gene and one in the waaO gene, whose products are involved in the synthesis of the core oligosaccharide moiety of lipopolysaccharides. Besides producing truncated lipopolysaccharides and displaying hypersensitivity to novobiocin and sodium dodecyl sulfate (SDS), these mutants were also shown by urea-SDS-polyacrylamide gel electrophoresis analysis to have reduced amounts of porins in their outer membranes. Moreover, they showed a reduced degradation of p-nitrophenyl phosphate and an increased resistance to ampicillin, two indications of a decrease in outer membrane permeability for small hydrophilic solutes. Additionally, ompC and ompF knockout mutants displayed levels of tolerance to the lactoperoxidase system similar to those displayed by the waa mutants. These results suggest that mutations which reduce the porin-mediated outer membrane permeability for small hydrophilic molecules lead to increased tolerance to the lactoperoxidase enzyme system because of a reduced uptake of OSCN.  相似文献   

8.
In order to study mechanisms underlying selective enzyme release from human leukocytes during phagocytosis, the effects were studied of compounds which affect microtubule integrity or the accumulation of cyclic nucleotides. Human leukocytes selectively extrude lysosomal enzymes (β-glucuronidase) from viable cells during phagocytosis of zymosan or immune complexes, or upon encounter with immune complexes dispersed along a non-phagocytosable surface such as a millipore filter. In each circumstance, lysosomal enzyme release was reduced by previous treatment of cells with pharmacological doses of drugs which disrupt microtubules (e.g. 10-3–10-5 M colchicine) or with agents which affect accumulation of adenosine 3'5'-monophosphate (cAMP) (e.g. 10-3 M cyclic nucleotides and 2.8 x 10-4–2.8 x 10-6 M prostaglandin E (PGE) and A (PGA) compounds). Preincubation of cells with 5 µg/ml cytochalasin B resulted in complete inhibition of zymosan ingestion, but not of adherence of zymosan particles to plasma membranes or selective enzyme release. In this system, in which enzyme release was independent of particle uptake, preincubation of cells with colchicine, vinblastine, dibutyryl cAMP, or PGE1 also reduced extrusion of lysosomal enzymes. When cell suspensions were incubated with membrane-lytic crystals of monosodium urate (MSU), cytoplasmic as well as lysosomal enzymes were released with subsequent death of the cells. However, enzyme release followed phagocytosis of crystals (as measured by enhanced C-1 oxidation of glucose) and was due to "perforation from within" of the lysosomal membrane, rather than lysis by crystals of the plasma membrane. Enzyme release after MSU ingestion was also reduced when cells were treated with pharmacological doses of the test agents. When cells were killed by Triton X-100, acting on the plasma membrane, C-1 oxidation of glucose was abolished and enzyme release could not be inhibited pharmacologically. These observations suggest that lysosomal enzyme release from human phagocytes can be an active process which accompanies plasma membrane stimulation, is independent of cell death, and may be controlled by cyclic nucleotides and agents which affect microtubules.  相似文献   

9.
This study investigated the sites of urate synthesis and catabolism in the gecarcinid land crab Gecarcoidea natalis by assaying spongy connective tissue, midgut gland, muscle and gill for xanthine oxidoreductase, the last enzyme involved in urate synthesis, and uricase and urease, the first and last enzymes involved in urate catabolism. The spongy connective tissue and midgut gland of the G. natalis contained activities of xanthine oxidoreductase and were considered to be sites of urate synthesis. The midgut gland had a high activity of xanthine oxidoreductase [(58.87±4.6 (SE) nmol urate produced g-1 wet wt. tissue min-1], 2.7 times the xanthine oxidoreductase activity contained within the spongy connective tissue, and was thought to be the main site of urate synthesis. Xanthine dehydrogenase (EC 1.1.1.204) was the only form of xanthine oxidoreductase detected within the tissues. Its presence means that the cost of synthesising urate de novo is relatively small (between 1 and 3 ATP). Uricase (EC 1.7.3.3) and urease (EC 3.5.1.5) activities were present in the tissues of G. natalis. Spongy connective tissue contained the highest activities of uricase [48.44±4.29 (SE) nmol urate consumed g-1 wet wt. tissue min-1] while the highest activities of urease [365.31±37.21 (SE) nmol urate consumed g-1 wet wt tissue min-1] were contained within the gills. From this evidence it is clear that G. natalis possesses the uricolytic pathway and hence the ability to catabolise urate, and urate catabolism is begun at the site of urate storage, the spongy connective tissue, and is completed at the gills. As the gills are the site of ammonia excretion in this species the ammonia produced from the catabolism of urate is probably excreted. The urate deposits within the body of G. natalis may be involved in temporary storage of nitrogenous wastes.  相似文献   

10.
Effect of adrenaline on 32P incorporation into rat fat-cell phospholipids   总被引:3,自引:3,他引:0  
1. The phospholipid composition of fat-cells prepared from rat epididymal fat-pad was determined. 2. The incorporation of [32P]Pi into the phospholipids of fat-cells incubated in glucose-free medium and the effect of adrenaline and of α- and β-adrenergic blocking agents, were studied. 3. Incorporation of [32P]Pi into fat-cell phospholipid increased with time; incubation with adrenaline resulted in increased incorporation that was related to the concentration of adrenaline. 4. The pattern of incorporation of [32P]Pi into the individual phospholipids of fat-cells after incubation for 1h was determined; adrenaline (5.4μm) resulted in increased incorporation into phosphatidylcholine. 5. Incubation of fat-cells with propranolol (34μm) and adrenaline (5.4μm) resulted in abolition of adrenaline-stimulated lipolysis; there was a decrease in the specific radioactivity of phosphatidylcholine and an increase in the specific radioactivity of phosphatidylethanolamine, phosphatidic acid, phosphatidylinositol and cardiolipin compared with cells incubated with adrenaline alone. 6. Incubation of fat-cells with phenoxybenzamine (0.1mm) and adrenaline (5.4μm) resulted in stimulation of lipolysis, and in diminished specific radioactivities of phosphatidylcholine, phosphatidic acid, phosphatidylinositol, phosphatidylglycerol and choline plasmalogen compared with cells stimulated with adrenaline alone.  相似文献   

11.
The addition of nitro blue tetrazolium (NBT) into the reaction of adrenaline autooxidation allows direct identification of superoxide anion formation (O2−⊙) as well as demonstration of kinetics of their accumulation in this superoxide-generating system. The kinetics of adrenochrome and O2−⊙ formation has been compared under the same conditions. Three possible approaches to the use of the adrenaline autooxidation reaction for the determination of superoxide dismutase activity (SOD) and revealing antioxidant properties of various compounds are discussed. Two of these approaches have been described previously: the spectro-photometric method of registration of adrenochrome, an end product of adrenaline autooxidation, at 347 nm (Sirota, 1999) and the polarographic method, which measures oxygen consumption used for O2−⊙ formation (Sirota, 2011). Here, a novel approach to this problem is presented; it is based on spectrophotometric determination of O2−⊙ using NBT. The employment of this approach results in a significant decrease of the pH value of carbonate buffer from 10.5 to 9.7 and a 4-fold decrease in the amounts of added adrenaline, thus creating milder conditions for the revealing and investigation of antioxidant properties of materials being examined.  相似文献   

12.
IntroductionBoth sugar-sweetened beverage (SSB) intake and body mass index (BMI) are associated with elevated serum urate concentrations and gout risk. The aim of this study was to determine whether the associations of SSB intake with serum urate and gout are moderated by BMI.MethodThe effects of chronic SSB intake on serum urate and gout status were analysed in a large cross-sectional population study. The effects of an acute fructose load on serum urate and fractional excretion of uric acid (FEUA) were examined over 180 minutes in a short-term intervention study. In all analyses, the responses were compared in those with BMI <25 mg/kg2 (low BMI) and ≥25 mg/kg2 (high BMI).ResultsIn the serum urate analysis (n = 12,870), chronic SSB intake was associated with increased serum urate in the high BMI group, but not in the low BMI group (Pdifference = 3.6 × 10−3). In the gout analysis (n = 2578), chronic high SSB intake was associated with gout in the high BMI group, but not in the low BMI group (Pdifference = 0.012). In the acute fructose loading study (n = 76), serum urate was increased in the high BMI group at baseline and throughout the observation period (PBMI group <0.0001), but there were similar acute serum urate increases in both BMI groups in response to the fructose load (Pinteraction = 0.99). The baseline FEUA was similar between the two BMI groups. However, following the fructose load, FEUA responses in the BMI groups differed (Pinteraction <0.0001), with increased FEUA at 120 minutes and 180 minutes in the low BMI group and reduced FEUA at 60 minutes in the high BMI group.ConclusionsThese data suggest that BMI influences serum urate and gout risk in response to chronic SSB intake, and renal tubular uric acid handling in response to an acute fructose load. In addition to many other health benefits, avoidance of SSBs may be particularly important in those with overweight/obesity to prevent hyperuricaemia and reduce gout risk.

Trials registration

Australian Clinical Trials Registry ACTRN12610001036000. Registered 24 November 2010.  相似文献   

13.
We previously reported that high micromolar concentrations of nitric oxide were able to oxidize mitochondrial cytochrome c at physiological pH, producing nitroxyl anion (Sharpe and Cooper, 1998 Biochem. J. 332, 9–19). However, the subsequent re-evaluation of the redox potential of the NO/NO- couple suggests that this reaction is thermodynamically unfavored. We now show that the oxidation is oxygen-concentration dependent and non stoichiometric. We conclude that the effect is due to an oxidant species produced during the aerobic decay of nitric oxide to nitrite and nitrate. The species is most probably nitrogen dioxide, NO2? a well-known biologically active oxidant. A simple kinetic model of NO autoxidation is able to explain the extent of cytochrome c oxidation assuming a rate constant of 3 × 106 M-1 s-1 for the reaction of NO2? with ferrocytochrome c. The importance of NO2? was confirmed by the addition of scavengers such as urate and ferrocyanide. These convert NO2? into products (urate radical and ferricyanide) that rapidly oxidize cytochrome c and hence greatly enhance the extent of oxidation observed. The present study does not support the previous hypothesis that NO and cytochrome c can generate appreciable amounts of nitroxyl ions (NO- or HNO) or of peroxynitrite.  相似文献   

14.
Cultures of Clostridium formicoaceticum and C. thermoaceticum growing on fructose and glucose, respectively, were shown to rapidly oxidize CO to CO2. Rates up to 0.4 μmol min−1 mg of wet cells−1 were observed. Carbon monoxide oxidation by cell suspensions was found (i) to be dependent on pyruvate, (ii) to be inhibited by alkyl halides and arsenate, and (iii) to stimulate CO2 reduction to acetate. Cell extracts catalyzed the oxidation of carbon monoxide with methyl viologen at specific rates up to 10 μmol min−1 mg of protein−1 (35°C, pH 7.2). Nicotinamide adenine dinucleotide, nicotinamide adenine dinucleotide phosphate and ferredoxin from C. pasteurianum were ineffective as electron acceptors. The catalytic mechanism of carbon monoxide oxidation was “ping-pong,” indicating that the enzyme catalyzing carbon monoxide oxidation can be present in an oxidized and a reduced form. The oxidized form was shown to react reversibly with cyanide, and the reduced form was shown to react reversibly with alkyl halides: cyanide inactivated the enzyme only in the absence of carbon monoxide, and alkyl halides inactivated it only in the presence of carbon monoxide. Extracts inactivated by alkyl halides were reactivated by photolysis. The findings are interpreted to indicate that carbon monoxide oxidation in the two bacteria is catalyzed by a corrinoid enzyme and that in vivo the reaction is coupled with the reduction of CO2 to acetate. Cultures of C. acidi-urici and C. cylindrosporum growing on hypoxanthine were found not to oxidize CO, indicating that clostridia mediating a corrinoid-independent total synthesis of acetate from CO2 do not possess a CO-oxidizing system.  相似文献   

15.
Lactoperoxidase catalyzes the oxidation of thiocyanate by hydrogen peroxide and an intermediary product is formed with antibacterial properties. The components of this system, with the exception of hydrogen peroxide, are present in milk. H2O2 may be introduced by means of enzymatic, generation and thus make the system complete. A two-enzyme system consisting of β–galactosidase and glucose oxidase has been developed for this purpose. The coupled enzyme reaction is shown to work with high efficiency at the neutral pH of milk although the enzymes as such, particularly lactases suitable for immobilization, have optimal activities at much lower pH values. The results indicate that the lactoperoxidase system may in this way be employed to inactivate bacteria present in milk.  相似文献   

16.
Uricase (urate: oxygen oxidoreductase; EC 1.7.3.3) from the rust Puccinia recondita was purified to electrophoretic homogeneity. Preparations with a specific activity of 8.4 U/mg were used for characterization of the enzyme, which showed a strong similarity to other plant and fungal urate oxidases. The enzyme had a pH optimum of 9.0, a K m of 35 μM for urate, and it was inhibited only by oxonate and xanthine. A molecular mass of 152 kDa was estimated for the native protein. SDS-PAGE analysis revealed a striking difference to most urate oxidases, since two different-sized subunits were detected. These results suggest that P. recondita uricase is a tetramer with two types of subunits. Received: 21 February 2001 / Accepted: 30 July 2001  相似文献   

17.
A membrane-bound NADH oxidase of an anaerobic alkaliphile, M-12 (a strain of Amphibacillus sp.), was solubilized with decanoyl N-methylglucamide and purified by chromatography on DEAE-Sepharose and hydroxyapatite. The purified enzyme appears to consist of a single polypeptide component with an apparent molecular mass of 56 kDa. The enzyme catalyzed the oxidation of NADH with the formation of H2O2 and exhibited a specific activity of 46 μmol NADH min–1 (mg protein)–1. NADPH did not serve as a substrate for the enzyme. The K m for NADH was estimated to be 0.05 mM. The enzyme exhibited a pH dependence for activity, with a pH optimum at approximately 9.5. The enzyme required a high concentration of salt and exhibited maximum activity in the presence of 600 mM NaCl. Received: 3 August 1998 / Accepted: 23 December 1998  相似文献   

18.
Rat renal proximal tubule cell membranes have been reported to contain uricase-like proteins that function as electrogenic urate transporters. Although uricase, per se, has only been detected within peroxisomes in rat liver (where it functions as an oxidative enzyme) this protein has been shown to function as a urate transport protein when inserted into liposomes. Since both the uricase-like renal protein and hepatic uricase can transport urate, reconstitution studies were performed to further characterize the mechanism by which uricase may function as a transport protein. Ion channel activity was evaluated in planar lipid bilayers before and after fusion of uricase-containing proteoliposomes. In the presence of symmetrical solutions of urate and KCl, but absence of uricase, no current was generated when the voltage was ramped between ±100 mV. Following fusion of uricase with the bilayer, single channel activity was evident: the reconstituted channel rectified with a mean slope conductance of 8 pS, displayed voltage sensitivity, and demonstrated a marked selectivity for urate relative to K+ and Cl. The channel was more selective to oxonate, an inhibitor of both enzymatic uricase activity and urate transport, than urate and it was equally selective to urate and pyrazinoate, an inhibitor of urate transport. With time, pyrazinoate blocked both its own movement and the movement of urate through the channel. Channel activity was also blocked by the IgG fraction of a polyclonal antibody to affinity purified pig liver uricase. These studies demonstrate that a highly selective, voltage dependent organic anion channel is formed when a purified preparation of uricase is reconstituted in lipid bilayers.This work was supported in part by the G. Harold and Leila Y. Mathers Charitable Foundation (E.L.P. and R.D.L.), the Irma T. Hirschl Trust (R.D.L.), National Institutes of Health grant DK08419 (B.A.K.) and a Grant-in-Aid from the American Heart Association, N.Y.C. Affiliate (R.G.A.).  相似文献   

19.
The gene xylBADP1 from Acinetobacter baylyi ADP1 (gene annotation number ACIAD1578), coding for a putative aryl alcohol dehydrogenase, was heterologously expressed in Escherichia coli BL21(DE3). The respective aryl alcohol dehydrogenase was purified by fast protein liquid chromatography to apparent electrophoretic homogeneity. The predicted molecular weight of 39,500 per subunit was confirmed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. According to the native Mw as determined by gel filtration, the enzyme forms dimers and therefore seems to be XylB related. The enzyme showed the highest activity at 40°C. For both the reduction and the oxidation reactions, the pH for optimum activity was 6.5. The enzyme was NADH dependent and able to reduce medium- to long-chain n-alkylaldehydes, methyl-branched aldehydes, and aromatic aldehydes, with benzaldehyde yielding the highest activity. The oxidation reaction with the corresponding alcohols showed only 2.2% of the reduction activity, with coniferyl alcohol yielding the highest activity. Maximum activities for the reduction and the oxidation reaction were 104.5 and 2.3 U mg−1 of protein, respectively. The enzyme activity was affected by low concentrations of Ag+ and Hg2+ and high concentrations of Cu2+, Zn2+, and Fe2+. The gene xylBADP1 seems to be expressed constitutively and an involvement in coniferyl alcohol degradation is suggested. However, the enzyme is most probably not involved in the degradation of benzyl alcohol, anisalcohol, salicyl alcohol, vanillyl alcohol, cinnamyl alcohol, or aliphatic and isoprenoid alcohols.  相似文献   

20.
In vitro effects of ketamine and bupivacaine drugs on bovine lactoperoxidase (LPO; E.C. 1.11.1.7) enzyme activity were investigated. Lactoperoxidase was purified with Amberlite CG 50 resin, CM Sephadex C-50 ion-exchange chromatography, and Sephadex G-100 gel filtration chromatography from skimmed bovine milk. Rz(A412/A280) value for the purified LPO was found to be 0.8. Inhibition or activation effects of the drugs on LPO enzyme were determined using 2,21-azino-bis (3-ethylbenzthiazoline-6 sulfonic acid) diammonium salt (ABTS) as a chromogenic substrate at pH = 6.0. The I50 values of ketamine and bupivacaine were 0.29 mM and 0.155 mM, respectively and the Ki constants for ketamine and bupivacaine were 0.019 ± 0.031 and 0.015 ± 0.021 mM, respectively; they were non-competitive inhibitors.  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