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.
N-Acetyltransferase (NAT), responsible for bioactivation and detoxification of arylamines, has been demonstrated to be widely distributed in many organisms ranging from humans to microorganisms. Using high performance liquid chromatography (HPLC) to analyze NAT activity in bacteria, the authors found that Pseudomonas aeruginosa exhibited high NAT activity with 2-aminofluorene (2-AF) as substrate. Characteristics of this bacterial NAT were further investigated. The N-acetylation catalyzed by this enzyme is an acetyl coenzyme A (AcCoA)-dependent reaction. As the concentration of AcCoA in the reaction mixture was increased, the apparent K(m) and Vmax for 2-AF increased. The K(m) and Vmax were 0.504 +/- 0.056 mM and 31.92 +/- 3.23 nmol/min/mg protein, respectively, for the acetylation of 2-AF with 0.5 mM AcCoA. The optimum pH for the enzyme activity was estimated to be around 8.5. It was active at a temperature range from 5 degrees C to 55 degrees C, with maximum activity at 37 degrees C. The enzyme activity was inhibited by divalent metal ions including Cu++, Fe++, Zn++, Ca++, Co++, Mn++, and Mg++, suggesting that a sulfhydryl group is involved in the N-acetylation activity. The three chemical modification agents, iodoacetamide, phenylglyoxal, and diethylpyrocarbonate, all exhibited a dose-, time-, and temperature-dependent inhibition effect. Preincubation of the NAT with AcCoA provided significant protection against the inhibition of iodoacetamide and diethylpyrocarbonate, but only partial protection against the inhibition of phenylglyoxal. These results indicate that cysteine, histidine, and arginine residues are essential for this bacterial enzyme activity, and the first two are likely to reside on the AcCoA binding site, but arginine residue may be located only near the AcCoA binding site. Our data demonstrate that P. aeruginosa possesses highly active N-acetyltransferase which shares a similar catalytic mechanism as that of higher organisms. These findings are very helpful for further investigating the role of arylamine NAT in this bacterial species.  相似文献   

2.
Arylamine N-acetyltransferase (NAT) activities with p-aminobenzoic acid and 2-aminofluorene were determined in Helicobacter pylori, a gram-negative rod bacteria collected from peptic ulcer patients. The NAT activity was determined using a acetyl CoA recycling assay and HPLC. Cytosols or suspensions of H. pylori with and without selected concentrations of emodin co-treatment showed different percentages of 2-aminofluorene and p-aminobenzoic acid acetylation. The data indicate that there were decreased NAT activity associated with increased emodin in H. pylori cytosols. As 400 microns of emodin can obviously inhibit NAT activity both in vitro and in vivo (inhibition rate 90% and 93% for 2-aminofluorene and p-aminobenzoic acid in vitro, and 90% and 92%, respectively, for both substrate in vivo). For in vitro examination, the apparent values of Km and Vmax were 3.12 +/- 0.38 mM and 15.20 +/- 3.16 nmol/min/mg protein for 2-aminofluorene, and 0.56 +/- 0.12 mM and 0.74 +/- 0.09 nmol/min mg protein for p-aminobenzoic acid. However, when emodin was added to the reaction mixtures, the values of apparent Km and Vmax were 2.40 +/- 0.32 mM and 10.62 +/- 0.04 nmol/min/mg protein for 2-aminofluorene, and 0.23 +/- 0.02 mM and 0.62 +/- 0.08 nmol/min/mg protein for p-aminobenzoic acid. For in vivo examination, the apparent Km and Vmax were 0.82 +/- 0.18 mM and 0.92 +/- 0.21 nmol/min/10 x 10(10) colony forming units (CFU) for 2-aminofluorene, and 0.78 +/- 0.14 mM and 0.52 +/- 0.06 nmol/min/ 10 x 10(10) (CFU) for p-aminobenzoic acid. However, when emodin was added to the reaction mixtures, the values of apparent Km and Vmax were 0.50 +/- 0.08 mM and 0.62 +/- 0.22 nmol/min/ 10 x 10(10) (CFU) for 2-aminofluorene, and 0.52 +/- 0.21 mM and 0.26 +/- 0.04 nmol/min/ 10 x 10(10) (CFU) for p-aminobenzoic acid. This report is the first finding of emodin inhibition of arylamine N-acetyltransferase activity in a strain of H. pylori.  相似文献   

3.
Earlier investigations from our laboratory demonstrated that human term placental peroxidase (HTPP) is capable of metabolism of xenobiotics and endogenous compounds. In this study, purified HTPP was found to bioactivate 2-aminofluorene (2-AF) in the presence of H2O2. 2-AF oxidation was studied spectrophotometrically while radiometry was employed to assess the bioactivation. The rate of oxidation and covalent binding to protein and DNA was dependent upon the pH of the reaction medium and the concentration of 2-AF, the enzyme, and H2O2. To observe maximal enzyme velocity of oxidation, the presence of 16.5 microM H2O2, 100 microM 2-AF, 37 micrograms of the enzyme protein/ml, and pH 7.2 was required. Under optimal assay conditions, the range of specific activity between 130 and 165 nmol of 2-AF oxidized/min/mg HTPP was observed. Using similar assay conditions, the magnitude of covalent binding of [3H]-2-AF to protein (BSA) and calf thymus DNA was found to be about 508 pmol bound/min/mg HTPP/mg BSA and 84 pmol bound/min/mg HTPP/mg DNA, respectively. Potassium cyanide and sodium azide, the known inhibitors of different peroxidases, significantly blocked both the oxidation and covalent binding of 2-AF in a dose dependent manner. These results strongly suggest that peroxidase may be one of the important pathways responsible for the bioactivation of arylamines in human term placenta.  相似文献   

4.
The role of two common polymorphisms of enzymes involved in the metabolism of drugs and carcinogens was studied in relation to prostate cancer. The gene encoding one of these enzymes (NAT2) is located in an area where frequent allelic loss occurs in prostate cancer. Mutations at the genes CYP2D6 and NAT2 were analysed by allele-specific polymerase chain reaction and restriction mapping in DNA from 94 subjects with prostate cancer and 160 male healthy control subjects. Eleven prostate specimens were analysed for genotype and enzymatic activities NAT2, CYP2D6 and CYP3A by using the enzyme-specific substrates sulphamethazine and dextromethorphan. Enzyme activities with substrate specificities corresponding to NAT2, CYP2D6 and CYP3A are present in human prostate tissue, with mean +/-s.d. activities of 4.8+/-4.4 pmol min(-1) mg(-1) protein, 156+/-91 and 112+/-72 nmol min(-1) mg(-1) protein respectively. The Km values for the prostate CYP2D6 and CYP3A enzyme activities corresponded to that of liver CYP2D6 and CYP3A activities, and the CYP2D6 enzyme activity is related to the CYP2D6 genotype. The N-acetyltransferase, in contrast, had a higher Km than NAT2 and was independent of the NAT2 genotype. The CYP2D6 and CYP3A enzymes, and an N-acetyltransferase activity that is independent of the regulation of the NAT2 gene, are expressed in human prostate tissue. The presence of carcinogen-metabolizing enzymes in human prostate with a high interindividual variability may be involved in the regulation of local levels of carcinogens and mutagens and may underlie interindividual differences in cancer susceptibility.  相似文献   

5.
Glutathione content and glutathione S-transferase enzyme activity as well as isoenzyme composition were studied in normal gastric cardia, normal squamous esophageal epithelium and corresponding malignant tumor of 10 patients with esophageal cancer. Mean values of glutathione (38 +/- 6 versus 36 +/- 12 nmol/mg protein) and glutathione S-transferase activity (532 +/- 44 versus 532 +/- 108 nmol/min mg protein) did not differ significantly between normal esophageal and tumor tissue. However, great individual differences exist. In two patients, glutathione S-transferase activity was much higher in the tumor (1081 and 1381 nmol/min mg protein) due to overexpression of class alpha, mu and pig glutathione S-transferases in one case, and of class mu and pi in the other case. In the other patients, glutathione S-transferase activity was equal (one case) or lower (seven cases) in the tumor. In normal gastric cardia glutathione content as well as glutathione S-transferase activity was significantly lower as compared to normal esophageal epithelium. In conclusion, in contrast to other gastrointestinal tumors, glutathione S-transferases are overexpressed in esophageal tumors in only a limited number of patients.  相似文献   

6.
The microsomal ethanol oxidizing system comprises an ethanol-inducible cytochrome P-4502E1, but the involvement of other P-450s has also been suggested. In our study, human CYP2E1, CYP1A2, and CYP3A4 were heterologously expressed in HepG2 cells, and their ethanol oxidation was assessed using a corresponding selective inhibitor: all three P-450 isoenzymes metabolized ethanol. Selective inhibitors-4-methylpyrazole (CYP2E1), furafylline (CYP1A2), and troleandomycin (CYP3A4)-also decreased microsomal ethanol oxidation in the livers of 18 organ donors. The P-450-dependent ethanol oxidizing activities correlated significantly with those of the specific monooxygenases and the immunochemically determined microsomal content of the respective P-450. The mean CYP2E1-dependent ethanol oxidation in human liver microsomes [1.41+/-0.11 nmol min(-1) (mg protein)(-1)] was twice that of CYP1A2 (0.61+/-0.07) or CYP3A4 (0.73+/-0.11) (p < 0.05). Furthermore, CYP2E1 had the highest (p < 0.05) specific activity [28+/-2 nmol min(-1) (nmol CYP2E1)(-1) versus 17+/-3 nmol min(-1) (nmol CYP1A2)(-1), and 12+/-2 nmol min(-1) (CYP3A4)(-1), respectively]. Thus, in human liver microsomes, CYP2E1 plays the major role. However, CYP1A2 and CYP3A4 contribute significantly to microsomal ethanol oxidation and may, therefore, also be involved in the pathogenesis of alcoholic liver disease.  相似文献   

7.
Genes for the 290 amino acid, 33-34 kDa cytosolic acetyltransferases (NAT1* and NAT2*) from rat and hamster were cloned and expressed in Escherichia coli. Active clones were selected by a simple visual test for their ability to decolorize 4-aminoazobenzene in bacterial medium by acetylation. These recombinant acetyltransferases were analyzed for: (i) N-acetyltransferase, which was assayed by the rate of acetyl coenzyme A-dependent N-acetylation of 2-aminofluorene (2-AF) or 4-aminoazobenzene (AAB); (ii) arylhydroxamic acid acyltransferase, assayed by N,O-acyltransfer with N-hydroxy-N-acetyl-2-aminofluorene. Both NAT2s showed first order increases in N-acetylation rates with increasing 2-AF or AAB concentrations between 5 and 100 microM, with apparent K(m) values of 22-32 and 62-138 microM respectively. Although under the same conditions the N-acetylation rates for the two NAT1s declined by > 50%, below 5 microM 2-AF or AAB, the NAT rate data fit Michaelis-Menten kinetics, and the apparent K(m) values were 0.2-0.9 microM. For N,O-acyltransferase, the apparent K(m) values of the NAT1s were approximately 6 microM, while the K(m) values of the NAT2s were approximately 20- to 70-fold higher. SDS-PAGE/Western blot analysis of the recombinant acetyltransferases gave apparent relative molecular weights (MWr) of approximately 31 kDa for both NAT1s and rat NAT2 and approximately 29 kDa for hamster NAT2. Comparable MWr values were observed for native hamster liver NAT1 and NAT2 and for rat NAT1 under the same conditions. Although we did not detect NAT2-like activity in rat liver cytosol previously, the present data show that the rat NAT2* gene does code for a functional acetyltransferase, with properties similar to those of hamster liver NAT2. The data also indicate that at low substrate concentrations, NAT1 would apparently play the predominant role in vivo in N-acetylation and N,O-acyltransfer of aromatic amine derivatives, including their metabolic activation to DNA-reactive agents.  相似文献   

8.
The present study was designed to measure angiotensin-converting enzyme (ACE) activity in the human ovary and in serum and to relate this activity to age, serum estradiol levels, and uterine and endometrial pathology. ACE activity was determined in 56 females by a radiometric assay using [3H]hippuryl-glycyl-glycine as substrate. Ovarian ACE activity, but not serum ACE, was found to increase with age (P < 0.01) and was significantly greater in postmenopausal subjects (n = 31; 1.35 +/- 0.05 nmol/mg.min) than in subjects with active ovaries (n = 21; 0.65 +/- 0.2 nmol/mg.min; P = 0.0033). Ovarian ACE activities in fertile women in the preovulatory phase (n = 14) and the postovulatory phase (n = 7) were not statistically different (0.66 +/- 0.23 and 0.63 +/- 0.17 nmol/mg.min, respectively). Serum ACE activities were similar in females with active and nonactive ovaries (87.6 +/- 5.0 vs. 81.7 +/- 5.3 nmol/mL-min, respectively). Serum estradiol levels in fertile women were significantly higher than those in postmenopausal women (P = 0.0023). Serum estradiol levels were negatively correlated with age (r = -0.46; P = 0.0041) and were not correlated with either serum ACE activity (r = 0.080; P = NS) or ovarian ACE activity. In summary, human ovarian ACE activity, but not serum ACE, is positively correlated with age. Serum estradiol levels decrease with age, but are not correlated with either ovarian or serum ACE activity. Endogenous serum estradiol levels had no apparent effect on ovarian or serum ACE activity. The presence of uterine pathology affects ovarian ACE activity. The cause of the increased ovarian ACE activity is not clear, but may be related to the aging process.  相似文献   

9.
Phospholipase D (E.C. 3.1.4.4.) was detected in isolated bovine rod outer segments (ROS) and its properties determined. The enzyme activity was assayed using either a sonicated microdispersion of 1,2-diacyl-sn-[2(3)H]glycerol-3-phosphocholine (PC), or [14C]ethanol. Using [3H]PC and ethanol as a substrate, we were able to detect the hydrolytic properties as well as the transphosphatidylation reaction catalyzed by phospholipase D (PLD): formation of [3H]phosphatidic acid and phosphatidylethanol [3H]PtdEt; whereas with [14C]ethanol or [3H]glycerol in the absence of exogenous PC, only transphosphatidylation reactions were detected (formation of [14C]PtdEt or [3H]phosphatidylglycerol, respectively). The use of varying concentrations of [3H]PC and 400 mM of ethanol gave an apparent Km value for PC of 0.51 mM and a Vmax value of 111 nmol x h(-1) x (mg protein)(-1). The activity was linear up to 60 min of incubation and up to 0.2 mg of protein. The optimal ethanol concentration was determined to be 400 mM, with an apparent Km of 202 mM and a Vmax value for ethanol of 125 nmol x h(-1) x (mg protein)(-1). A clear pH optimum was observed around 7. PLD activity was increased in the presence of 3-[(3-cholamidopropyl)dimethylammonio]-1-propane-sulfonate or sodium deoxycholate and inhibited with Triton X-100. The enzyme activity was also activated in the presence of Ca2+ or Mg2+ (1 mM) although these ions were not required for measuring PLD activity. The high specific activity of PLD found in purified ROS compared to the activity found in other subcellular fractions of the bovine retina suggests that this enzymatic activity is native to ROS. The present report is the first evidence of PLD activity associated with photoreceptor ROS.  相似文献   

10.
Uptake of norcholansulfonate (3 alpha, 7 alpha, 12 alpha-trihydroxy-24-nor-5 beta-cholan-23-sulfonate), an isogeometric analogue of cholate into isolated rat liver hepatocytes occurs only by saturable transport. In order to identify the transport systems involved, uptake of norcholansulfonate was studied using 7 beta-NBD-NCT ({N-[7-(4-nitrobenzo-2-oxa-1,3-diazol)]-7 beta-amino-3 alpha,12 alpha-dihydroxy-5 beta-cholan-24-oyl})-2'-aminoethanesulfonate) as a competing substrate. For transport of both bile salt derivatives, which mutually inhibit their mediated transport competitively, the existence of at least three transport systems must be assumed. Uptake studies using the cloned hepatic Na+/cholyltaurine cotransporting polypeptide stably expressed in CHO cells (Chinese hamster ovary cells) showed that both bile salt derivatives were transported and furnished the definite KT values of this single transport system and the ratio of the maximal uptake velocities. On the basis of these data, uptake of both bile salt derivatives into rat hepatocytes and their mutual competitive inhibition could be analyzed for three transport systems. The maximal flux rates J2 and the half-saturation constants KT2 in the presence of Na+ (143 mM) are for norcholansulfonate: J1(Na+ 143) = 1.0 +/- 0.2 nmol/(min . mg protein), KT1(Na+ 143) = 15 +/- 4 microM, J2(Na+ 143) = 0.5 +/- 0.2 nmol/(min.mg protein), KT2(Na+ 143) = 15 +/- 2 microM, J3(Na+ 143) = 0.5 +/- 0.2 nmol/(min.mg protein), KT3(Na+ 143) = 60 +/- 15 microM, and for 7 beta-NBD-NCT J1(Na+ 143) = 0.14 +/- 0.04 nmol/(min.mg protein), KT1(Na+ 143) = 3.1 +/- 0.5 microM, J2(Na+ 143) = 0.014 +/- 0.005 nmol/(min.mg protein), KT2(Na+ 143) = 21 +/- 2 microM, J3(Na+ 143) = 1.0 +/- 0.1 nmol/(min.mg protein), KT3(Na+ 143) = 190 +/- 25 microM. The kinetic parameters are in accordance with the assumptions that the cloned Na+/cholyltaurine cotransporting polypeptide represents transport system 2 and that the kinetically identified additional transport system 1 is either strictly or partially Na(+)-dependent.  相似文献   

11.
The role of different cytochrome P450 isozymes (CYP) in the N-demethylation of chlorimipramine and chlorpromazine has been investigated in liver microsomes from rats by studying the effects of multiple subchronic doses of chlorimipramine, chlorpromazine, phenobarbital and beta-naphthoflavone on the N-demethylation of ethylmorphine, mono-N-demethyl-chlorimipramine and chlorpromazine and on the hydroxylation of aniline. With control microsomes, CYP-dependent metabolism of chlorimipramine and chlorpromazine (100 nmol; 30 min incubation) resulted in the formation of predominantly chlorimipramine (46.5 +/- 4.9 nmol) whereas chlorpromazine (14.1 +/- 0.9 nmol) accounted for only part of the overall metabolism of chlorpromazine. Multiple doses of chlorimipramine increased the capacity of microsomes to N-demethylate ethylmorphine (9.8 +/- 0.73 and 6.08 +/- 0.06 nmol min(-1) (mg protein)(-1) for chlorimipramine-treated and control rats, respectively) as well as itself (4.65 +/- 0.25 and 3.10 +/- 0.33 nmol min(-1) (mg protein)(-1), respectively). Multiple doses of chlorpromazine induced aniline-hydroxylase activity (1.11 +/- 0.16 and 0.94 +/- 0.06 nmol min(-1) (mg protein)(-1) for chlorimipramine and control microsomes, respectively) but the capacity to N-demethylate itself was unchanged. Phenobarbital treatment induced ethylmorphine N-demethylation activity, but did not affect N-demethylation activity, towards chlorimipramine and chlorpromazine. In control microsomes the N-demethylation capacity of chlorimipramine or chlorpromazine (0.160 +/- 0.025 and 0.015 +/- 0.003 nmol min(-1) (mg protein)(-1), respectively) was one order of magnitude lower than that of chlorimipramine or chlorpromazine. The capacity to N-demethylate either chlorimipramine or chlorpromazine was increased by treatment with either phenobarbital or beta-naphthoflavone. In control microsomes, sulphaphenazole markedly inhibited both chlorimipramine-N-mono- and di-N-demethylation, whereas quinidine markedly inhibited the rate of formation of chlorpromazine. The CYP2C and CYP2D subfamilies seem to be involved in the mono N-demethylation of chlorimipramine and chlorpromazine, respectively. Moreover the CYP1A and CYP2B subfamilies might participate in the N-demethylation of either chlorimipramine or chlorpromazine. This could have important implications in the clinical use of chlorimipramine and chlorpromazine in view of the genetic polymorphism of CYP2C and CYP2D isozymes in man.  相似文献   

12.
Mice constitutively express glutathione S-transferase mGSTA3-3 in liver. This isoform possesses uniquely high conjugating activity toward aflatoxin B1-8,9-epoxide (AFBO), thereby protecting mice from aflatoxin B1-induced hepatocarcinogenicity. In contrast, rats constitutively express a closely related GST isoenzyme, rGSTA3-3, with low AFBO activity and, therefore, are sensitive to aflatoxin B1 exposure. Although the two GSTs share 86% sequence identity and have similar catalytic activities toward 1-chloro-2,4-dinitrobenzene (CDNB), they have an approximately 1000-fold difference in catalytic activity toward AFBO. To identify amino acids that confer high activity toward AFBO, non-conserved rGSTA3-3 residues were replaced with mGSTA3-3 residues in two regions believed to form the substrate binding site. Twenty-one mutant rGSTA3-3 enzymes were generated by site-directed mutagenesis using combinations of nine different residues. Except for the E208D mutant, single mutations of rGSTA3-3 produced enzymes with no detectable AFBO activity. Generally, AFBO conjugation activity increased in additive fashion as mGSTA3-3 residues were introduced into the rGSTA3-3 enzyme with the six site mutant E104I/H108Y/Y111H/L207F/E208D/V217K displaying the highest AFBO activity (40 nmol/mg/min) of all the mutant enzymes. When this mutant enzyme was further modified by three additional substitutions (D103E/I105M/V106I) AFBO conjugation activity decreased 14-fold to 2. 8 nmol/mg/min. Although wild-type mGSTA3-3 AFBO conjugation activity (265 nmol/mg/min) could not be obtained by our rGSTA3-3 mutants, we were able to identify six mGSTA3-3 residues; Ile104, Tyr108, His111, Phe207, Asp208, and Lys217 that, when collectively substituted into rGSTA3-3, substantially increased (>200-fold) glutathione conjugation activity toward AFBO.  相似文献   

13.
Renal proximal tubules isolated from the rat possess nitric oxide synthase (NOS) activity that is calcium/calmodulin dependent and stereoselectively inhibited by NG-monomethyl-arginine (NMMA). In the absence of added Ca2+ and calmodulin, activity was reduced 84 +/- 13% compared with the activity in the presence of 2 mM Ca2+ and 25 micrograms/mL calmodulin. Inhibition by EGTA (10 mM) was 95 +/- 5% and by calmidazolium (R24571, 250 microM) was 99 +/- 1%. Inhibition by L-NMMA (100 microM) was 78 +/- 13% and by D-NMMA (100 microM) was 7 +/- 7%. The majority of NOS activity was found in the soluble fraction. NOS activity in isolated proximal tubules was also examined 6 hr after a single i.v. injection of lipopolysaccharide. Activity was increased significantly (P < 0.05) in the soluble fraction by 2-fold [from 0.320 +/- 0.052 to 0.648 +/- 0.046 (nmol/mg protein/30 min)] and in the particulate fraction by 3-fold [from 0.081 +/- 0.030 to 0.256 +/- 0.034 (nmol/mg protein/30 min)]. All activities were inhibited by EGTA. These data demonstrate that proximal tubules express a calcium/calmodulin-dependent NOS activity that is increased in vivo by lipopolysaccharide.  相似文献   

14.
Anandamide (N-Arachidonoylethanolamine) amidohydrolase catalyzing hydrolysis of anandamide was characterized in mice. The enzymatic activity was highest in the liver, followed by the brain and testis. Negligible activity was found in heart, lung and spleen. The activity in brain and liver was mainly localized in the microsomal fractions. Kinetic experiments demonstrated that Km (microM) and Vmax (nmol/min/mg protein) for the brain microsomes were 9.3 and 2.58, respectively, while those for the hepatic microsomes were 180 and 18.9, respectively. The activity in the microsomes from the liver and brain was markedly inhibited by Cu2+, Hg2+, Se4+, phenylmethylsulfonylfluoride and sodium dodecylsulfate. Brain but not hepatic microsomal enzyme activity was inhibited by delta9-tetrahydrocannabinol, cannabidiol and cannabinol. Kinetic parameters demonstrated that the inhibition by the cannabinoids was competitive in nature. Relatively high distribution of the enzyme activity in brain suggests an importance of the enzyme in the central nervous system to regulate the neuromodulatory fatty-acid amides.  相似文献   

15.
V79 (Chinese hamster lung fibroblast) cell lines expressing a functional recombinant phenobarbital-inducible rat liver UDP-glucuronosyltransferase (UGT), i.e., UGT2B1, were established. Western blot analysis of positive colonies, using anti-rat liver UGT antibodies, revealed the presence of an immunoreactive polypeptide of the expected molecular mass of 52 kDa. The substrate specificity of the recombinant enzyme toward > 100 compounds was determined. Phenolic and alcoholic substrates included 4-methylumbelliferone, 4-hydroxybiphenyl, chloramphenicol, and testosterone, but a range of carboxylic acids of both endogenous (medium-chain saturated fatty acids, long-chain polyunsaturated fatty acids, and bile acids) and exogenous (profen nonsteroidal anti-inflammatory drugs, fibrate hypolipidemic agents, and sodium valproate) origin were also accepted, indicating that the enzyme was capable of forming both ether- and ester-type glucuronides from various structurally unrelated compounds. Determination of apparent kinetic constants for the glucuronidation by UGT2B1 of selected aglycones revealed a high maximal velocity toward the 3-position of morphine (49.3 +/- 2.2 nmol/min/mg of protein), compared with other known substrates such as 4-methylumbelliferone (2.67 +/- 0.11 nmol/min/mg of protein) or clofibric acid (0.06 +/- 0.02 nmol/min/mg of protein). To gain a better insight into the mechanisms underlying the apparently wide substrate specificity of UGT2B1, series of structurally related compounds were tested as potential substrates. The rate of glucuronidation of unbranched saturated fatty acids and omega,omega,omega-triphenylalkanoic acids increased progressively with increasing alkyl chain length and then declined, with the best substrates in these two homologous series being decanoic acid and 4,4,4-triphenylbutanoic acid, respectively. Glucuronidation of para-substituted phenols always proceeded at a higher rate than that of the corresponding para-substituted benzoic acids. This could mean that the aglycon hydroxyl group was better positioned in the enzyme active site in the case of phenols. Alternatively, if the initial interaction with the enzyme required the aglycon to be in the protonated uncharged form, then the observation could be explained by the difference in ionization between phenols and benzoic acids at the incubation pH used. The introduction of a bulky alkyl group into the para-position led to increases of up to 300-fold in the rate of glucuronidation, probably as a result of the increased aglycon lipophilicity. Finally, the enzyme showed a degree of stereo- and regiospecificity, preferring (S)-ibuprofen to the R-enantiomer (Vmax/Km, 3.06 and 1.10 microliters/min/mg of protein, respectively) and glucuronidating lithocholic acid but not hyodeoxycholic acid, which differs by only a single hydroxyl group.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
This study was designed to assess the effects of aspirin on arylamine N-acetyltransferase (NAT) activities in the bacterium Klebsiella pneumoniae using high performance liquid chromatography to measure the acetylation of 2-aminofluorene (2-AF) with or without aspirin. Cytosols or suspensions of K. pneumoniae with or without specific concentrations of aspirin co-treatment showed different percentages of 2-AF acetylation. The data indicated that there was decreased NAT activity associated with increased levels of aspirin in K. pneumoniae cytosols and in intact bacteria. For the cytosol examination, the apparent values of Km and Vmax decreased 0.59- and 0.58-fold after co-treated with 40 microM aspirin, respectively, for 2-AF. For the intact bacteria examination, the apparent values of Km and Vmax decreased 0.60- and 0.67-fold after co-treated with 40 microM aspirin, respectively, for 2-AF. This report is the first demonstration to show that aspirin can decrease N-acetyltransferase activity in the bacterium K. pneumoniae.  相似文献   

17.
Endothelial nitric oxide synthase (NOS) is a constitutively expressed flavin-containing heme protein that catalyzes the formation of NO from L-arginine, NADPH, and molecular oxygen. We purified bovine endothelial NOS from transfected embryonic kidney cells by conventional chromatographic techniques and characterized the activity of the detergent-solubilized enzyme. Endothelial NOS displays a much lower specific activity of NO synthesis (143 +/- 11 nmol NO/min/mg enzyme) than the constitutive neuronal NOS or inducible NOS isoforms. Like the neuronal isoform, endothelial NOS requires binding of Ca2+/calmodulin to achieve Vmax NO synthase activity; however, we observed a basal level of NO synthesis even when Ca2+/calmodulin was omitted and 0.5 mM EDTA was present in the assay solution. Moreover, endothelial NOS demonstrates a high-affinity bonding interaction with calmodulin such that the enzyme as purified has a NO synthase activity at about 80% of Vmax. We also observed a more than twofold increase in NADPH consumption by endothelial NOS when it was coupled to arginine oxygenation as opposed to when oxygen is activated in the absence of substrate. Substrate binding was also shown to stimulate heme reduction in the absence of added calmodulin. Thus, the enzymatic synthesis of NO from L-arginine by endothelial NOS appears to be partially regulated by binding of both calmodulin and substrate. These findings for endothelial NOS represent a significant departure from the enzymatic properties of the other constitutive NOS isoform, neuronal NOS, and we interpret this result in terms of the physiological implications.  相似文献   

18.
AIMS: To investigate, in healthy volunteers, the relationships between the plasma concentrations (C, ng ml(-1)) of zabiciprilat, the active metabolite of the angiotensin I-converting enzyme inhibitor (ACEI) zabicipril, and the effects (E) induced on plasma converting enzyme activity (PCEA, nmol ml(-1) min(-1)), brachial and femoral artery flows (BAF, FAF, ml min(-1)), and brachial and femoral vascular resistances (BVR, FVR, mmHg x s ml(-1)) after a single oral administration of two doses (0.5 and 2.5 mg) of zabicipril. METHODS: The study was placebo-controlled, randomized, double-blind and crossover. E was related to C by the Hill model, E = Emax x Cgamma/(CE50gamma + Cgamma), fitted to the data of both doses simultaneously. RESULTS: We obtained (mean +/- s.d.) Emax = -99 +/- 1%, CE50 = 2.2 +/- 1.0 ng ml(-1) and gamma = 1.0 +/- 0.4 for PCEA, Emax = 55 +/- 26 ml min(-1), CE50 = 5.1 +/- 4.0 ng ml(-1) and gamma = 2.4 +/- 1.6 for BAF, and Emax = -45 +/- 10%, CE50 = 2.0 +/- 1.3 ng ml(-1) and gamma = 2.3 +/- 1.4 for BVR. The parameters obtained for FAF and FVR were similar to those obtained for BAF and BVR, respectively. The CE95 (C required to induce 95% of Emax) varies from 7 to 17 ng ml(-1) for haemodynamic effects. CONCLUSIONS: As zabiciprilat peak plasma concentrations average 20 ng ml(-1) after the 2.5 mg dose of zabicipril, this dose of the drug should be sufficient to induce optimal haemodynamic effects.  相似文献   

19.
Arylamine N-acctyltransferase (NAT) activities with p-aminobenzoic acid (PABA) and 2-aminofluorene (2-AF) were determined in the bacterium Helicobacter pylori collected from peptic ulcer patients. Two assay systems were performed, one with cellular cytosols, the other with intact cell suspensions. Cytosols or suspensions of H. pylori with or without specific concentrations of diallyl sulfide (DAS) or diallyl disulfide (DADS) co-treatment showed different percentages of 2-AF and PABA acetylation. The data indicated that there was decreased NAT activity associated with increased levels of DAS or DADS in H. pylori cytosols and suspensions. Viability studies on H. pylori demonstrated that DAS or DADS elicited dose-dependent bactericide affects on H. pylori cultures. The data also indicated that DAS and DADS decreased the apparent values of K(m) and Vmax of NAT enzyme from H. pylori in both systems examined. This report is the first demonstration that garlic components can affect H. pylori growth and NAT activity.  相似文献   

20.
The enzyme system responsible for the N-deacetylation of eprinomectin in rats was characterized. Tissue and subcellular studies showed that the hydrolysis activity was localized mainly in liver microsomes. Apparent KM and Vmax values calculated from Lineweaver-Burk plots were 53 microM and 0.81 nmol/mg/min for male rats and 70 microM and 4.99 nmol/mg/min for female rats, respectively. Pretreatment of male rats with dexamethasone, phenobarbital, and pregnenolone 16alpha-carbonitrile increased the activity by more than 3-fold. Paraoxon and bis-4-nitrophenylphosphate strongly inhibited the deacetylase activity at concentrations as low as 1 microM. The hydrolysis activity also was inhibited by SKF525, but less effectively. Eserine strongly inhibited the activity at 1 x 10(-4) M. HgCl2 decreased the activity to about 40% at a concentration of 1 x 10(-4) M. FeCl3, CaCl2, MgCl2, and EDTA had little effect on the hydrolysis of eprinomectin, whereas NaF slightly increased the activity to 118%. Thus, the inhibition study suggested that eprinomectin deacetylase resembled "B" type carboxylesterase/amidases. The hydrolysis activity of eprinomectin and isocarboxazid, a specific substrate of RL2 [Hosokawa, M, Maki T and Satoh T (1987) Mol Pharmacol 31:579-584], by liver microsomes from rats treated with various cytochrome P-450 inducers correlated well (r = 0.92). Also, elusion profiles of esterase by gel filtration and ion exchange chromatography demonstrated that the active protein(s) for eprinomectin and isocarboxazid hydrolysis coeluted. Thus, RL2 or an enzyme system similar to RL2 is responsible for the N-deacetylation of eprinomectin.  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