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 750 毫秒
1.
Purified bovine liver beta-glucuronidase (beta-D-glucuronide glucuronohydrolase, EC 3.2.1.32) and wheat germ acid phosphatase (orthophosphoric monoesterphosphohydrolase, EC 3.1.3.2) were inhibited with freshly dissolved and 24 h aquated tetrahaloaurate (III) compounds. Rate and equilibrium inhibition constants were measured. From this data two acid phosphatases species were observed. Equilibrium inhibition constants ranged from 1 to 12.5 microM for the various gold compounds toward both enzymes. The first order rate constants ranged between 0.005 and 0.04 min.-1 for most reactions with the exception of the fast reacting acid phosphatase which had values as high as 2.6 and 2.8 min.-1. It is observed that the beta-glucuronidase is rapidly inhibited during the equilibrium phase before the more slower reaction covalent bond formation takes place. The acid phosphatases form the covalent bonds more rapidly, especially the faster reacting species suggesting a unique difference in the active site geometry to that of the more slowly reacting species. The tightly bonded gold (III)-enzyme complex is probably the reason for its toxicity and non-anti-inflammatory use as a drug.  相似文献   

2.
Fumarase (EC 4.2.1.2) and mitochondrial L-malate dehydrogenase (EC 1.1.1.37) were both inhibited by NaAuCl4 and KAuBr4. The inhibition for both was measured as a function of gold complex concentration and aquation time, and the NaAuCl4 inhibition was also measured in the presence of 0.15 M NaCl. Regeneration of the enzyme activity after NaAuCl4 inhibition using L-cysteine, L-methionine and NaCN was also investigated. Sodium dodecyl sulfate (SDS) acrylamide gel electrophoresis and amino acid analysis was performed on the NaAuCl4 inhibited enzymes as well as on ribonuclease A (EC 3.1.26.2), lysozyme (EC 3.2.1.17) and liver alcohol dehydrogenase (EC 1.1.1.1). It was observed that the inhibition was proportional to the gold complex concentration but decreased markedly after aquation of the complex. In the presence of NaCl the initial rate of inactivation is essentially unaffected unless the complex has been aquated and then the initial rate is increased. Gel electrophoresis on gold complex-enzyme mixtures show polymerization for ribonuclease and lysozyme and amino acid analysis indicates that no oxidation has taken place. From these results, a binding mechanism is postulated for the inhibition of the dehydrogenases by direct displacement of a halide ligand, probably by two groups on the enzyme, at least one of which may be a sulfur containing acid.  相似文献   

3.
Gold(I) thioglucose in the presence of excess glutathione (GSH) leads to strong and reversible inhibition of selenium-glutathione peroxidase (EC 1.11.19) around neutral pH. Binding at equilibrium and competition studies demonstrated that the most reduced form of the active site selenocysteine is the only binding site for gold(I) Steady-state kinetics indicate that gold(I) forms a dead-end complex with glutathione peroxidase in competition with the reduction of hydroperoxide. The apparent K1 is 2 3 μM at pH 7.6,37°C and 1 mM GSH. Theoretical models of inhibition were assessed by the use of linear least-squares fitting to a generalized integrated rate equation. The results are consistent with trapping of gold(I) at the active site in the form of a mixed bidentate selenolato-thiolate complex involving GSH and the active site selenocysteine. The kinetics of inhibition imply that the resting form of glutathione peroxidase in the presence of excess GSH is also within the enzyme cycle. This rules out the existence of selenium(+lV) species in the redox cycle of the active site when t-butylhydroperoxide is used as a substrate. Electronic properties of selenium and gold as well as a large relief of inhibition by selenocysteine suggest that a very stable interaction should be obtained between Se(-II) and gold(I) through covalent bonding. These results suggest that glutathione peroxidase could be a target of gold drugs used in the treatment of rheumatoid arthritis.  相似文献   

4.
The efflux of [14C]citrate from rat heart mitochondria was significantly greater with L-malate as the extramitochondrial substrate as compared with [12C]citrate, isocitrate or phosphoenolpyruvate. The concentration of L-malate required for half-maximal rate of efflux of citrate was 0.45 mM and the maximum velocity was 0.36 nmol min-1 mg-1 mitochondrial protein at 23 degrees C. This citrate transporter was inhibited by 1,2,3-benzenetricarboxylate and palmitoyl-CoA but not to the same extent as these compounds inhibit the tricarboxylate carrier in rat liver mitochondria. The apparent inability of these mitochondria to transport citrate in the inward direction necessitates the presence of a cytosolic citrate removal pathway. We propose that the enzymes of this pathway in rat heart could be ATP citrate (pro-3S)-lyase (EC 4.1.3.a) and carnitine acetyltransferase (EC 2.3.1.7), both of which we demonstrate to have adequate activity in both the fed and fasted state. An hypothesis has been put forward to account for the inhibition of rat heart phosphofructokinase by citrate in the fasted state incorporating these properties of the citrate transporter and ATP citrate (pro-3S)-lyase.  相似文献   

5.
Rat liver malate dehydrogenase (decarboxylating) (NADP) ((L-malate: NADP) oxidoreductase (oxaloacetate-decarboxylating), EC 1.1.1.40) was purified and crystallized from medium containing 30 mM Tris-HCl buffer (pH 7.7), 5 mM MgCl2 and 2 mM 2-mercaptoethanol. The enzyme formed rhomboid crystals free from coenzyme, and appeared homogeneous on isoelectric focusing. The crystalline enzyme had an isoelectric point of pH 6.3. Amino acid analysis showed that it contained more acidic amino acids than basic ones.  相似文献   

6.
Pigeon liver malic enzyme (malate dehydrogenase (decarboxylating), EC 1.1.1.40) was reversibly inactivated by periodate-oxidized NADP in a biphasic manner. The reversibility could be made irreversible by treating the modified enzyme with sodium borohydride. The inactivation showed saturation kinetics and could be prevented by nucleotide (NADP or NADPH). Fully protection was afforded by the combination of NADP, Mn2+ and L-malate. Oxidized NADP was also found to be a coenzyme and noncompetitive inhibitor of L-malate in the oxidative decarboxylase reaction catalyzed by malic enzyme.  相似文献   

7.
Microsomes from liver or kidney of untreated rainbow trout hydroxylated lauric acid specifically at the (omega-1) position. Turnover numbers for liver (2.72 min-1) and kidney (14.1 min-1) were decreased seven- and twofold, respectively, following treatment with beta-naphthoflavone. Laurate hydroxylation activity from untreated trout hepatic microsomes was sensitive to inhibition by SKF-525A, but was not sensitive to metyrapone and only partially inhibited by alpha-naphthoflavone. The temperature optimum of laurate (omega-1) hydroxylation in trout liver microsomes was 25-30 degrees C. The Km and Vmax for (omega-1)- hydroxylaurate formation was 50 microM and 1.63 nmol min-1 mg-1, respectively, in liver and 20 microM and 3.95 nmol min-1 mg-1, respectively, in kidney from untreated trout microsomes. (omega-1) Hydroxylation of laurate, in both liver and kidney microsomes, was sensitive to an antibody raised against a previously purified cytochrome P-450 isozyme (LM2) of trout liver microsomes, which has been shown to be active towards aflatoxin B1. Antibody to the major isozyme of cytochrome P-450 ( LM4b , active towards benzo(a)pyrene) induced by beta-naphthoflavone did not inhibit (omega-1) hydroxylation of laurate in microsomes from untreated or beta-naphthoflavone-treated trout.  相似文献   

8.
NADP-linked malic enzyme [EC 1.1.1.40] was highly purified from Escherichia coli W cells. The purified enzyme was homogeneous as judged by ultracentrifugation and gel electrophoresis. The apparent molecular weights obtained by sedimentation equilibrium analysis, from diffusion and sedimentation constants, and by disc electrophoresis at various gel concentrations were 471,000, 438,000, and 495,000, respectively. The subunit molecular weights obtained by sedimentation equilibrium analysis in the presence of 6 M guanidine hydrochloride and gel electrophoresis in the presence of sodium dodecyl sulfate were 76,000 and 82,000, respectively. The sedimentation coefficient (S(0)20, W) was 13.8S, and the molecular activity was 44,700 min-1 at 30 degrees C. The amino acid composition of the enzyme was determined, and the results were compared with those of NAD-linked malic enzyme from the same organism and those of pigeon liver NADP-linked malic enzyme. The partial specific volume was calculated to be 0.738 ml/g. The Km value for L-malate was 2.3 mM at pH 7.4. Malonate, tartronate, glutarate, and DL-tartrate competitively inhibited the activity. The saturation profile for L-malate exhibited a marked cooperativity in the presence of both chloride ions and acetyl-CoA. However, acetyl-CoA alone did not show cooperativity or produce inhibition in the absence of chloride ions. Vmax and Km were determined as a function of pH. The optimum pH for the reaction was 7.8. Inspection of the Dixon plots suggested that three ionizable groups of the enzyme are essential for the enzyme activity. In addition to the oxidative decarboxylase activity, the enzyme preparation exhibited divalent metal ion-dependent oxaloacetate decarboxylase and alpha-keto acid reductase activities. Based on the above results, the molecular properties of the enzymatic reaction are discussed.  相似文献   

9.
In the present study we have used beef heart submitochondrial preparations (BH-SMP) to demonstrate that a component of mitochondrial Complex I, probably the NADH dehydrogenase flavin, is the mitochondrial site of anthracycline reduction. During forward electron transport, the anthracyclines doxorubicin (Adriamycin) and daunorubicin acted as one-electron acceptors for BH-SMP (i.e. were reduced to semiquinone radical species) only when NADH was used as substrate; succinate and ascorbate were without effect. Inhibitor experiments (rotenone, amytal, piericidin A) indicated that the anthracycline reduction site lies on the substrate side of ubiquinone. Doxorubicin and daunorubicin semiquinone radicals were readily detected by ESR spectroscopy. Doxorubicin and daunorubicin semiquinone radicals (g congruent to 2.004, signal width congruent to 4.5 G) reacted avidly with molecular oxygen, presumably to produce O2-, to complete the redox cycle. The identification of Complex I as the site of anthracycline reduction was confirmed by studies of ATP-energized reverse electron transport using succinate or ascorbate as substrates, in the presence of antimycin A or KCN respiratory blocks. Doxorubicin and daunorubicin inhibited the reduction of NAD+ to NADH during reverse electron transport. Furthermore, during reverse electron transport in the absence of added NAD+, doxorubicin and daunorubicin addition caused oxygen consumption due to reduction of molecular oxygen (to O2-) by the anthracycline semiquinone radicals. With succinate as electron source both thenoyltrifluoroacetone (an inhibitor of Complex II) and rotenone blocked oxygen consumption, but with ascorbate as electron source only rotenone was an effective inhibitor. NADH oxidation by doxorubicin during BH-SMP forward electron transport had a KM of 99 microM and a Vmax of 30 nmol X min-1 X mg-1 (at pH 7.4 and 23 degrees C); values for daunorubicin were 71 microM and 37 nmol X min-1 X mg-1. Oxygen consumption at pH 7.2 and 37 degrees C exhibited KM values of 65 microM for doxorubicin and 47 microM for daunorubicin, and Vmax values of 116 nmol X min-1 X mg-1 for doxorubicin and 114 nmol X min-1 X mg-1 for daunorubicin. In marked contrast with these results, 5-iminodaunodrubicin (a new anthracycline with diminished cardiotoxic potential) exhibited little or no tendency to undergo reduction, or to redox cycle with BH-SMP. Redox cycling of anthracyclines by mitochondrial NADH dehydrogenase is shown, in the accompanying paper (Doroshow, J. H., and Davies, K. J. A. (1986) J. Biol. Chem. 261, 3068-3074), to generate O2-, H2O2, and OH which may underlie the cardiotoxicity of these antitumor agents.  相似文献   

10.
An NAD(P)-dependent 3 alpha-hydroxysteroid dehydrogenase (EC 1.1.1.50) was purified to homogeneity from rat liver cytosol, where it is responsible for most if not all of the capacity for the oxidation of androsterone, 1-acenaphthenol and benzenedihydrodiol (trans-1,2-dihydroxycyclohexa-3,5-diene). The dehydrogenase has many properties (substrate specificity, pI, Mr, amino acid composition) in common with the dihydrodiol dehydrogenase (EC 1.3.1.20) purified from the same source [Vogel, Bentley, Platt & Oesch (1980) J. Biol. Chem. 255, 9621-9625]. Since 3 alpha-hydroxysteroids are by far the most efficient substrates, the enzyme is more appropriately designated a 3 alpha-hydroxysteroid dehydrogenase. It also promotes the NAD(P)H-dependent reductions of quinones (e.g. 9,10-phenanthrenequinone, 1,4-benzoquinone), aromatic aldehydes (4-nitrobenzaldehyde) and aromatic ketones (4-nitroacetophenone). The dehydrogenase is not inhibited by dicoumarol, disulfiram, hexobarbital or pyrazole. The mechanism of the powerful inhibition of this enzyme by both non-steroidal and steroidal anti-inflammatory drugs [Penning & Talalay (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 4504-4508] was examined with several substrates. Most non-steroidal anti-inflammatory drugs are competitive inhibitors (e.g. Ki for indomethacin, 0.20 microM for 9,10-phenanthrenequinone reduction at pH 6.0, and 0.835 microM for androsterone oxidation at pH 7.0), except for salicylates, which act non-competitively (e.g. Ki for aspirin, 650 microM for androsterone oxidation). The inhibitory potency of these agents falls sharply as the pH is increased from 6 to 9. Most anti-inflammatory steroids are likewise competitive inhibitors, except for the most potent (betamethasone and dexamethasone), which act non-competitively. The enzyme is inhibited competitively by arachidonic acid and various prostaglandins.  相似文献   

11.
1. Subcellular fractionation of rat, guinea pig and human livers showed that aldehyde dehydrogenase metabolizing gamma-aminobutyraldehyde was exclusively localized in the cytoplasmic fraction in all three mammalian species. 2. Total gamma-aminobutyraldehyde activity of aldehyde dehydrogenase was found to be ca 0.41, 0.3 and 0.24 mumol NADH min-1 g-1 tissue, respectively in rat, guinea pig and human liver, with more than 95% of activity in the cytoplasm. 3. Partially purified cytoplasmic isozyme from rat liver showed similar chromatographic behavior and kinetic properties to the E3 isozyme isolated from human liver. 4. The rat isozyme was insensitive to disulfiram (40 microM) and to magnesium (160 microM) and had Km values of 5 microM (pH 7.4) for gamma-aminobutyraldehyde, 7.5 microM (pH 9.0) for propionaldehyde and 4 microM (pH 7.4) for NAD.  相似文献   

12.
The nature of the interaction between Escherichia coli cytidine deaminase and the phosphapyrimidine nucleoside 1 has been studied kinetically and spectrophotometrically. Compound 1 was designed as a transition-state analog, and is a potent, slow-binding inhibitor of cytidine deaminase (Ashley, G. W., and Bartlett, P. A. (1982) Biochem. Biophys. Res. Commun. 108, 1467-1474). We present evidence that the binding of 1 is reversible, with no covalent linkage between the enzyme and 1. At pH 6, the rate of recovery of enzyme activity from dissociation of the E X I complex is strongly dependent on the concentration of E X I, indicating that the inhibitor dissociates reversibly. UV difference spectroscopy reveals that the chromophore of 1 is unaltered on binding to the enzyme, thus eliminating the possibility of reversible, covalent modification of the enzyme. For the binding of the active beta-anomers of 1 to cytidine deaminase, the following kinetic parameters were determined at pH 6: kon = 8300 M-1 S-1, koff = 7.8 X 10(-6) S-1, Ki = 0.9 nM. We were also able to observe and characterize time-dependent inhibition of E. coli cytidine deaminase by tetrahydrouridine, 3. This interaction involves involves initial formation of a loose complex (KD = 1.2 microM), followed by isomerization in a slow step to give a more tightly bound complex (Ki = 0.24 microM) with forward and reverse rate constants kf = 3.81 min-1 and kr = 0.95 min-1, respectively.  相似文献   

13.
Yeast hexokinase is a homodimer consisting of two identical subunits. Yeast hexokinase was inactivated by 2-aminothiophenol at 25 degrees C (pH 9.1). The reaction followed pseudo-first-order kinetics until about 70% of the phosphotransferase activity was lost. About 0.65 mol of 2-aminothiophenol/mol of hexokinase was found to be bound after the 70% loss of the enzyme activity. Completely inactivated hexokinase showed a stoichiometry of about 1 mol of 2-aminothiophenol bound/mol of the enzyme. The evidence obtained from kinetic experiments, stoichiometry of the inactivation reaction and fluorescence emission measurements suggested site-site interaction (weak negative co-operativity) during the inactivation reaction. The approximate rate constants for the reversible binding of 2-aminothiophenol to the first subunit (KI) and for the rate of covalent bond formation with only one site occupied (k3) were 150 microM and 0.046 min-1 respectively. The inactivation reaction was pH-dependent. Dithiothreitol, 2-mercaptoethanol and cysteine restored the phosphotransferase activity of the hexokinase after inactivation by 2-aminothiophenol. Sugar substrates protected the enzyme from inactivation more than did the nucleotides. Thus it is concluded that the inactivation of the hexokinase by 2-aminothiophenol was a consequence of a covalent disulphide bond formation between the aminothiol and thiol function at or near the active site of the enzyme. Hexokinase that had been completely inactivated by 2-aminothiophenol reacted with o-phthalaldehyde. Fluorescence emission intensity of the incubation mixture containing 2-aminothiophenol-modified hexokinase and o-phthalaldehyde was one-half of that obtained from an incubation mixture containing hexokinase and o-phthalaldehyde under similar experimental conditions. The intensity and position of the fluorescence emission maximum of the 2-aminothiophenol-modified hexokinase were different from those of the native enzyme, indicating conformational change following modification. Whereas aliphatic aminothiols were completely ineffective, aromatic aminothiols were good inhibitors of the hexokinase. Cyclohexyl mercaptan weakly inhibited the enzyme. Inhibition of the hexokinase by heteroaromatic thiols was dependent on the nature of the heterocyclic ring and position of the thiol-thione equilibrium. The inhibitory function of a thiol is associated with the following structural characteristics: (a) the presence of an aromatic ring, (b) the presence of a free thiol function and (c) the presence of a free amino function in the close proximity of the thiol function.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
For the purpose of developing highly sensitive and convenient determination of plasmalogens, the high-performance liquid chromatography (HPLC) method using radioactive iodine ((125)I) was investigated. Radioactive triiodide (1-) ion ((125)I(3)(-)), which is an actual iodine form capable of reacting with vinyl ether bond ([bond]CH(2)[bond]O[bond]CH[double bond]CH[bond]) of plasmalogens, could be safely and efficiently produced by oxidizing a commercial radioactive sodium iodine (Na(125)I) with hydrogen peroxide (H(2)O(2)) under acid condition (pH 5.5-6.0), which is called iodine-125 reagent. I(3)(-) specifically reacted with plasmalogens at the molar ratio of 1:1 in methanol, and 1 or 2 mol of plasmalogens was involved in the binding with iodine per iodine atom, resulting in the formation of stable iodine-binding phospholipids. The HPLC system with Diol column and acetonitrile/water as a mobile phase was available for separating iodine-binding phospholipids from nonbinding free iodine and for separately eluting iodine-binding phospholipids derived from choline and ethanolamine plasmalogens. Using iodine-125 reagent (1.85 MBq/ml), plasmalogens were detectable at high sensitivity of 10,000-15,000 cpm/nmol, which is more than 1000-fold higher sensitivity than the classical determination with nonradioactive iodine. Plasmalogen concentrations in human plasma were measured with the HPLC system and determined as, on average, 129.1+/-31.3 microM (n=8) in a 1.2 content ratio of choline to ethanolamine plasmalogens, a concentration that nearly agrees with the value reported previously.  相似文献   

15.
L-Cysteinatogold(I) was prepared by the reaction of L-cysteine with KAuBr4 in acidic media and its solubility determined from pH 4 to 10. The solubility at pH 7.4 and 37 degrees C is 1 microM. In the presence of excess cysteine, the solubility increases because of formation of bis(L-cysteinato)gold(I). The equilibrium-constant for formation of the bis complex is 2.1 +/- 0.4 X 10(-3), which at pH 7.4 CORRESPONDs to an apparant formation constant of 4.4 X10(4). The formation of the bis adduct was confirmed by chromatographic separation of the products of the reaction between [35S]-L-cysteine and Na2AuTM. This complex elutes with Kav = 1.15 which allows it to be distinguished from other gold thiolates that might form in vivo. The bis(cysteinato)gold(I) complex is shown to be present in kidney cytosol isolated from rats given Na2AuTM in vivo. When additional cysteine is added to the cytosol in vitro, the peak at 1.15 is increased, but if glutathione is added, the low molecular weight gold elutes at Kav = 1.00, which is taken as evidence for the existence of bis(cysteinato)gold(I) in the cytosol preparation. The amount of gold present as bis(cysteinato)gold(I) after 4 different dose schedules has been measured and found to increase with the total cytosol gold concentration. L-Cysteinatogold(I) does not dissolve in the presence of bovine serum albumin to form an adduct.  相似文献   

16.
The specific binding of (3H) isoproterenol to isolated fat cells of human and rat was characterized. Binding of (3H) isoproterenol to isolated fat cells of rat was saturable with 420 pmol of (3H) isoproterenol bound/100 mg of lipid. Half-maximal saturation occurred at 5 microM providing an estimate of the equilibrium dissociation constant, KD, for the interaction of (3H) isoproterenol with its binding sites. Kinetic analysis of (3H) isoproterenol binding provided a value of 2.01 x 10(4) min-1. M-1 for the forward bimolecular rate constant, k1. Dissociation of (3H) isoproterenol was a first order reaction with a rate constant, k2, of 0.62 x 10(-1) min-1. The ratio k2/k1 = 3.07 microM provides an independent measurement of the KD for the interaction of (3H) isoproterenol with its binding sites which is in agreement with the values obtained by steady state analysis (3 to 5 microM). The apparent equilibrium dissociation constant, KD, for the interaction of (3H) isoproterenol with its receptor in human fat cells obtained by steady state analysis was 1 to 0.9 microM. Scatchard- and Hill-analysis suggest the possibility of different negatively cooperative interactions among the binding sites in human and rat. beta-Adrenergic agonists competed for the binding sites. The order of potency was isoproterenol greater than epinephrine greater than norepinephrine. Compounds such as DOPA, dopamine and (m-Hydroxyphenyl)2-methyl-aminoethanol which are structurally related to catecholamines had little or no affinity for (3H) isoproterenol binding sites.  相似文献   

17.
The 3 alpha-hydroxysteroid dehydrogenase (EC 1.1.1.50) of rat brain cytosol has been purified to apparent homogeneity. The purification procedure involves six successive steps, includes one affinity chromatography, and yields enzyme which displays a 1,550-fold enhancement in specific activity. The homogeneous enzyme has a Km of 8.0 microM for 5 alpha-dihydrotestosterone, a Vmax of 1.3 mumol of 3 alpha-androstanediol formed per h/mg of protein, and displays a preference for NADPH. It appears to be the major activity responsible for the reduction of 5 alpha-dihydrotestosterone in this tissue and may play a pivotal role in brain androgen metabolism. The homogeneous enzyme has several properties in common with the 3 alpha-hydroxysteroid dehydrogenase purified from rat liver cytosol (Penning, T. M., Mukharji, I., Barrows, S., and Talalay, P. (1984) Biochem. J. 222, 601-611). It is a monomer with a molecular weight of 31,000, it has a pI of 5.5, and it is potently inhibited by the nonsteroidal anti-inflammatory drugs (IC50 value for indomethacin = 2.0 microM). The potency of inhibition observed for the brain enzyme parallels that observed for cyclooxygenase: indomethacin greater than fenamates greater than l-methylpyrrole acetic acids greater than arylpropionic acids greater than salicylates greater than acetaminophen. Examination of a variety of steroidal contraceptives as modulators of the dehydrogenase indicates that ethinylestradiol is a very poor inhibitor (IC50 = 100 microM), while 6-medroxyprogesterone acetate (Provera) is an extremely potent inhibitor (IC50 = 0.2 microM). The possibility exists that brain androgen metabolism may be altered by the nonsteroidal anti-inflammatory drugs and synthetic progestins.  相似文献   

18.
A series of 9-substituted adenine derivatives inhibited adenylate cyclase activity (ATP pyrophosphate-lyase (cyclizing) EC 4.6.1.1) of a particulate preparation of human blood platelets. A 3--6 fold elevation of adenylate cyclase activity by prostaglandin E1 (PGE1) was inhibited in a concentration-related manner by 9-(tetrahydro-5-methyl-2-furyl) adenine (SQ 22,538), 9-(tetrahydro-2-furyl) adenine (SQ 22,536), 9-cyclopentyladenine (SQ 22,534), 9-furfuryladenine (sQ 4647) and 9-benzyladenine (SQ 218611). The I50 values ranged from 21 microM for SQ 22,538 to 140 microM for SQ 21,611. These same adenine derivatives reversed the inhibition by PGE1 of ADP-induced aggregation and the PGE1-stimulated elevation of adenosine 3':5'-monophosphate (cyclic AMP). The reversal of platelet aggregation inhibition by SQ 22,536 and SQ 4647 was concentration-related with I50 values of 30 microM in each case, whereas SQ 22,534 and SQ 21,611 reversed inhibition by 30% at 100 microM. SQ 22,536, SQ 22,534 and SQ 21,611 also blocked the increase in cyclic AMP levels in a concentration-related manner with I50 values of 1, 4 and 60 microM, respectively. SQ 4647 inhibited the elevation of cyclic AMP by more than 85% at 1000 microM. The adenine derivatives had no effect on platelet aggregation or on cyclic AMP levels in the absence of PGE1. These results provide additional evidence that the inhibition of platelet aggregation by PGE1 is mediated by cyclic AMP.  相似文献   

19.
1. Protein disulphide-isomerase (EC 5.3.4.1) and glutathione-insulin transhydrogenase (EC 1.8.4.2) were resolved by covalent chromatography. Both activities, in a partially purified preparation from bovine liver, bind covalently as mixed disulphides to activated thiopropyl-Sepharose 6B, in a new stepwise elution procedure protein disulphide-isomerase is displaced in mildly reducing conditions whereas glutathione-insulin transhydrogenase is only displaced by more extreme reducing conditions. 2. This together with evidence for partial resolution of the two activities by ion-exchange chromatography, conclusively establishes that the two activities are not alternative activities of a single bovine liver enzyme. 3. Protein disulphide-isomerase, partially purified by a published procedure, has now been further purified by covalent chromatography and ion-exchange chromatography. The final material is 560-fold purified relative to a bovine liver homogenate; it has barely detectable glutathione-insulin transhydrogenase activity. 4. The purified protein disulphide-isomerase shows a single major band on sodium dodecyl sulphate/polyacrylamide-gel electrophoresis corresponding to a mol.wt. of 57000. 5. The purified protein disulphide-isomerase has Km values for 'scrambled' ribonuclease and dithiothreitol of 23 microgram/ml and 5.4 microM respectively and has a sharp pH optimum at 7.5. The enzyme has a broad thiol-specificity, and several monothiols, at 1mM, can replace dithiothreitol. 6. The purified protein disulphide-isomerase is completely inactivated after incubation with a 2-3 fold molar excess of iodoacetate. The enzyme is also significantly inhibited by low concentrations of Cd2+ ions. These findings strongly suggest the existence of a vicinal dithiol group essential for enzyme activity. 7. When a range of thiols were used as co-substrates for protein disulphide-isomerase activity, the activities were found to co-purify quantitatively, implying the presence of a single protein disulphide-isomerase of broad thiol-specificity. Glutathione-disulphide transhydrogenase activities, assayed with a range of disulphide compounds, did not co-purify quantitatively.  相似文献   

20.
Activities of glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12) (GAP-DH) and aldolase (EC 4.1.2.13) in cells of Clostridium perfringens that had been inhibited with sodium nitrite were investigated. A complete loss in GAP-DH activity and a 67% decrease in aldolase activity were observed when growth of C. perfringens was inhibited. There was also a 91% decrease in the concentration of free sulfhydryl groups of soluble cellular components. Dithiothreitol restored some activity to inactive GAP-DH from sodium nitrite-inhibited cells, indicating that a loss of reduced sulfhydryl groups was involved in the inactivation of the enzyme. The evidence presented suggests that sodium nitrite inhibition of C. perfringens may involve an interaction of sodium nitrite as nitrous acid with sulfhydryl-containing constituents of the bacterial cell.  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