首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The glucose analog 5-thio-d-glucose, a potent inhibitor of glucose transport across membranes, was examined as an acceptor and/or inhibitor of lactose synthetase (UDP-galactose: D-glucose 1-galactosyltransferase, EC 2.4.1.22). Thioglucose was an effective acceptor for lactose synthetase with a Km of 7.4 mM. Under identical conditions the Km for D-glucose in this reaction was 5.4 mM. Thioglucose was 45 to 50% as effective an acceptor as D-glucose. Thioglucose acted as a pseudo substrate having a different Km and Vmax. Thus, thioglucose could be considered to the be a competitive substrate for lactose synthetase. thetase. The product of the lactose synthetase reaction with thioglucose as an acceptor had a thin-layer chromatographic retardation factor slightly higher than that for lactose. Upon treatment of the reaction product with β-galactosidase, galactose and thioglucose were released. These observations suggest that the product of the lactose synthetase reaction with thioglucose was thiolactose.  相似文献   

2.
Radioactive sucrose, supplied through the cut base to Pisum sativum epicotyls, was transported to the growing apex (plumule and hook) and used there for the synthesis mainly of uridine diphosphoglucose (UDP- glucose), fructose and cell wall glucan. Enzyme extracts of the apical tissue contained sucrose synthetase activity which was freely reversible, i.e. formed UDP-glucose and fructose from sucrose (pH optimum = 6·6 for the cleavage reaction, Km for sucrose = 63 mM). Particulate fractions of the same tissue contained a β-glucan synthetase which utilized UDP-glucose for formation of alkali-soluble and -insoluble products (pH optimum = 8·4, Km for UDP-glucose = 1·9 mM). Values for Vmax and yields of these two synthetase activities were sufficient to account for observed rates of cellulose deposition during epicotyl growth (15–25 μg/hr/epicotyl). When soluble pea enzyme was supplied with sucrose and UDP at pH 6·6 and then the preparation was supplemented with particles bearing β-glucan synthetase at pH 8·4, the glucose moiety of sucrose was converted to glucan in vitro. The results indicate that it is feasible for these synthetases to co-operate in vivo to generate β-glucan for expanding cell walls.  相似文献   

3.
The levels of cyclic 2,3-diphosphoglycerate (cDPG) in methanogenic bacteria are governed by the antagonistic activities of cDPG synthetase and cDPG hydrolase. In this paper we focus on the synthetase from Methanobacterium thermoautotrophicum. The cytoplasmic 150 kDa enzyme catalyzed cDPG synthesis from 2,3-diphosphoglycerate (apparent Km=21 mM), Mg2+ (Km=3.1 mM) and ATP (Km=1–2 mM). In batch-fed cultures, the enzyme was constitutively present (6–6.5 nmol per min per mg protein) during the different growth phases. In continuous cultures, activity decreased in response to phosphate limitation. The synthetase reaction proceeded with maximal rate at pH 6 and at 65° C and was specifically dependent on high (>0.3M) K+ concentrations. The reaction conditions remarkably contrasted to those of cDPG degradation catalyzed by the previously described membrane-bound cDPG hydrolase.Abbreviations cDPG Cyclic 2,3-diphosphoglycerate - 2,3-DPG 2,3-Diphosphoglycerate - 2-PG 2-Phosphoglycerate - 3-PG 3-Phosphoglycerate  相似文献   

4.
The initial reactions possibly involved in the acrobic and anaerobic metabolism of aromatic acids by a denitrifying Pseudomonas strain were studied. Several acyl CoA synthetases were found supporting the view that activation of several aromatic acids preceeds degradation. A benzoyl CoA synthetase activity (AMP forming) (apparent K m values of the enzyme from nitrate grown cells: 0.01 mM benzoate, 0.2 mM ATP, 0.2 mM coenzyme A) was present in aerobically grown and anaerobically, nitrate grown cells when benzoate or other aromatic acids were present. In addition to benzoate and fluorobenzoates, also 2-amino-benzoate was activated, albeit with unfavorable K m (0.5 mM 2-aminobenzoate). A 2-aminobenzoyl CoA synthetase (AMP forming) was induced both aerobically and anaerobically with 2-aminobenzoate as growth substrate which had a similar substrate spectrum but a low K m for 2-aminobenzoate (<0.02 mM). Anaerobic growth on 4-hydroxybenzoate induced a 4-hydroxybenzoyl CoA synthetase, and cyclohexanecarboxylate induced another synthetase. In contrast, 3-hydroxybenzoate and phenyl-acetate grown anaerobic cells appeared not to activate the respective substrates at sufficient rates. Contrary to an earlier report extracts from aerobic and anaerobic 2-aminobenzoate grown cells catalysed a 2-aminobenzoyl CoA-dependent NADH oxidation. This activity was 10–20 times higher in aerobic cells and appeared to be induced by 2-aminobenzoate and oxygen. In vitro, 2-aminobenzoyl CoA reduction was dependent on 2-aminobenzoyl CoA NAD(P)H, and oxygen. A novel mechanism of aerobic 2-aminobenzoate degradation is suggested, which proceeds via 2-aminobenzoyl CoA.  相似文献   

5.
Glutamine synthetase (EC 6.3.1.2) was purified to homogeneity from a free-living nitrogen fixing bacteria, Bacillus polymyxa. The holoenzyme, relative molecular mass (Mr) of 600 000 is composed of monomeric sub-units of 60 000 (Mr). The isoelectric point of the sub-units was 5.2. The pH optimum for the biosynthetic and transferase enzyme activity was 8.2 and 7.8, respectively. The apparent K m values (K m app ) in the biosynthetic reaction for glutamate, NH4Cl and ATP were 3.2, 0.22 and 1 mM, respectively. In the transferase reaction the K m values for glutamine, hydroxylamine and ADP were 6.5, 3.5 and 8×10-4 mM respectively. L-Methionine-D-L-sulfoximine was a very potent inhibitor in both biosynthetic and transferase reactions. Similar to most Gram positive bacteria there was no evidence of in vivo adenylylation and the enzyme seemed to be mainly regulated by feed-back mechanism.Abbreviations PMSF phenylmethylsulfonylfluoride - TCA trichloroacetic acid - GS glutamine synthetase - MSO L-Methionine-D-L-sulfoximine - SDS-PAGE sodium dodecyl sulfatepolyacrylamide gel electrophoresis - SVPDE snake venum phosphodiesterase  相似文献   

6.
l-Alanine dehydrogenase was found in extracts of the antibiotic producer Streptomyces clavuligerus. The enzyme was induced by ammonia, and the level of induction was dependend on the extracellular concentration. l-Alanine was the only amino acid able to induce alanine dehydrogenase. The enzyme was characterized from a 38-fold purified preparation. Pyruvate (K m =1.1 mM), ammonia (K m =20 mM) and NADH (K m =0.14 mM) were required for the reductive amination, and l-alanine (K m =9.1 mM) and NAD (K m =0.5 mM) for the oxidative deaminating reaction. The aminating reaction was inhibited by alanine, serine and NADPH. Alanine inhibited uncompetitively with respect to NADH (K i =1.6 mM) and noncompetitively with respect to ammonia (K i =2.0 mM) and pyruvate (K i =3.0 mM). In the aminating reaction 3-hydroxypyruvate, glyoxylate and 2-oxobutyrate could partially (6–7%) substitute pyruvate. Alanine dehydrogenase from S. clavuligerus differed with respect to its molecular weight (92000) and its kinetic properties from those described for other microorganisms.Abbreviation Alanine-DH l-alanine:NAD oxidoreductase  相似文献   

7.
Desulfobacter postgatei is an acetate-oxidizing, sulfate-reducing bacterium that metabolizes acetate via the citric acid cycle. The organism has been reported to contain a si-citrate synthase (EC 4.1.3.7) which is activated by AMP and inorganic phosphate. It is show now, that the enzyme mediating citrate formation is an ATP-citrate lyase (EC 4.1.3.8) rather than a citrate synthase. Cell extracts (160,000xg supernatant) catalyzed the conversion of oxaloacetate (apparent K m=0.2 mM), acetyl-CoA (app. K m=0.1 mM), ADP (app. K m=0.06 mM) and phosphate (app. K m=0.7 mM) to citrate, CoA and ATP with a specific activity of 0.3 mol·min-1·mg-1 protein. Per mol citrate formed 1 mol of ATP was generated. Cleavage of citrate (app. K m=0.05 mM; V max=1.2 mol · min-1 · mg-1 protein) was dependent on ATP (app. K m=0.4 mM) and CoA (app. K m=0.05 mM) and yielded oxaloacetate, acetyl-CoA, ADP, and phosphate as products in a stoichiometry of citrate:CoA:oxaloacetate:ADP=1:1:1:1. The use of an ATP-citrate lyase in the citric acid cycle enables D. postgatei to couple the oxidation of acetate to 2 CO2 with the net synthesis of ATP via substrate level phosphorylation.  相似文献   

8.
The Km for ammonia for glutamine synthetase and glutamate dehydrogenase was measured in enzyme extracts from Skeletonema costatum (Grev.) Cleve. At similar physiological pH and temperature the half-saturation constant for glutamine synthetase was 29 μM, whereas for GDH it was 28mM. On the basis of relative enzymic activity, as well as substrate affinity, it is suggested that glutamine synthetase is the enzyme primarily responsible for the incorporation of ammonium into the amino acid pool, when extracellular nitrogen is at ecological concentrations.  相似文献   

9.
K m for L-phenylalanine, L-glutamic acid, L-aspartic acid, and the corresponding keto acids were calculated, as well as V max was measured for the following pairs of substrates: L-phenylalanine-2-ketoglutarate, L-phenylalanine-oxaloacetate, L-glutamic acid-phenylpyruvate, and L-aspartic acid-phenylpyruvate for aminotransferases PAT1, PAT2, and PAT3 from Erwinia carotovora catalyzing transamination of phenylpyruvate. The ping-pong bi-bi mechanism was shown for the studied aminotransferases. The substrate inhibition (K s) of PAT3 with 2-ketoglutarate and oxaloacetate was 10.23 ± 3.20 and 3.73 ± 1.99 mM, respectively. It was shown that L-β-(N-benzylamino)alanine was a competitive inhibitor with respect to L-phenylalanine for PAT1 (K i = 0.32 ± 0.07 mM, K m = 0.45 ± 0.1 mM, V max = 11. 6 ± 0.4 U/mg) at 25 mM concentration of 2-ketoglutarate in the reaction medium. L-β-(N-methylamino)alanine is a noncompetitive inhibitor with respect to L-phenylalanine for PAT3 (K I = 138.4 ± 95.4 mM, K m = 13.7 ±3.9 mM, V max = 18.6 ± 4.1 U/mg) at 2 mM concentration of 2-ketoglutarate in the reaction medium. L-stereo isomers of nonprotein analogues of aromatic amino acids were studied as substrates for PAT1, PAT2, and PAT3. L-β-(2-Br-phenyl)alanine, L-β-(4-Br-phenyl)alanine, L-β-(2-F-phenyl)alanine, and L-(2-F)tryptophan were good substrates for all three aminotransferases; L-α-methyl-β-(2-Br-phenyl)alanine and L-O-benzyltyrosine were substrates only for PAT3; L-β-(4-F-phenyl)alanine was a substrate for PAT1 and PAT3. Thus, these analogues of aromatic amino acids can be stereoselectively synthesized using the studied aminotransferases in the presence of the corresponding keto acids.  相似文献   

10.
The estrogen synthetase present in human placental microsomes appears to be dependent on the cooperative interaction of the reduced cofactors NADPH and NADH for optimal activity. Using steady-state concentrations of either cofactor, it was found that while the estrogen synthetase activity followed hyperbolic saturation kinetics with NADPH (Kmapp = 14 μM), the enzyme followed sigmoidal saturation kinetics when the cofactor was NADH, with the half-maximum velocity attained at a cofactor concentration of 1.1 mm. The maximum velocity obtained with NADPH as the cofactor was greater than with corresponding concentrations of NADH. Estrogen synthetase activity in the presence of NADH was not due to NADPH contamination. NADH, in the presence of small concentrations of NADPH (0.5 to 5 μm), stimulated significantly the rate of estrogen formation from androstenedione by placental microsomes and, in addition, the enzyme saturation kinetics changed from sigmoidal to hyperbolic, thus mimicking the effect of NADPH. Estrogen synthetase activity, measured in the presence of 1 mm NADH, was stimulated in a dose-dependent manner by NADPH (Kmapp = 0.4 μM NADPH) and, when the enzyme was measured in the presence of 5 μm NADPH, the activity was stimulated in a dose-dependent manner by NADH (Kmapp = 45 μM NADH). Estrogen synthetase activity measured in the presence of NADH, without and with NADPH (1 μm) remained linear both with time of incubation for approximately 15 min and with microsomal protein concentration up to 3 mg/ml. The apparent Km of estrogen synthetase for androstenedione, when measured in the presence of NADH, was 1 μm. The synergistic interaction between NADH and NADPH in stimulating placental estrogen synthetase activity observed in vitro may, conceivably, take place in vivo in the intact placenta.  相似文献   

11.
S-adenosylmethionine synthetase was studied from bloodstream forms of Trypanosoma brucei brucei, the agent of African sleeping sickness. Two isoforms of the enzyme were evident from Eadie Hofstee and Hanes-Woolf plots of varying ATP or methionine concentrations. In the range 10–250 μM the Km for methionine was 20 μM, and this changed to 200 μM for the range 0.5–5.0 mM. In the range 10–250 μM the Km for ATP was 53 μM, and this changed to 1.75 mM for the range 0.5–5.0 mM. The trypanosome enzyme had a molecular weight of 145 kDa determined by agarose gel filtration. Methionine analogs including selenomethionine, L-2-amino-4-methoxy-cis but-3-enoic acid and ethionine acted as competitive inhibitors of methionine and as weak substrates when tested in the absence of methionine with [14C]ATP. The enzyme was not inducible in procyclic trypomastigotes in vitro, and the enzyme half-life was > 6 h. T. b. brucei AdoMet synthetase was inhibited by AdoMet (Ki 240 μM). The relative insensitivity of the trypanosome enzyme to control by product inhibition indicates it is markedly different from mammalian isoforms of the enzyme which are highly sensitive to AdoMet. Since trypanosomes treated with the ornithine decarboxylase antagonist DL-α-difluoromethylornithine accumulate AdoMet and dcAdoMet (final concentration ≈ 5 mM), this enzyme may be the critical drug target linking inhibition of polyamine synthesis to disruption of AdoMet metabolism.  相似文献   

12.
Summary. Branched-chain amino acid aminotransferase was purified by several column chromatographies from Helicobacter pylori NCTC 11637, and the N-terminal amino acid sequence was analyzed. The enzyme gene was sequenced based on a putative branched-chain amino acid aminotransferase gene, ilvE of H. pylori 26695, and the whole amino acid sequence was deduced from the nucleotide sequence. The enzyme existed in a homodimer with a calculated subunit molecular weight (MW) of 37,539 and an isoelectric point (pI) of 6.47. The enzyme showed high affinity to 2-oxoglutarate (K m = 0.085 mM) and L-isoleucine (K m = 0.34 mM), and V max was 27.3 μmol/min/mg. The best substrate was found to be L-isoleucine followed by L-leucine and L-valine. No activity was shown toward the D-enantiomers of these amino acids. The optimal pH and temperature were pH 8.0 and 37 °C, respectively.  相似文献   

13.
The ability of the enzyme subtilisin DY for the synthesis of derivatives of DL-aspartic acid which are differently N and C-terminal protected and semiproducts of the peptide synthesis was investigated. The enzyme reaction was characterized by high yields and a comparatively short reaction time. Two of the substrates, Z-D,L-Asp-(OMe)2 and PhAc-D,L-Asp-(OMe)2, were hydrolyzed for about 15 min; the reaction time for Boc-D,L-Asp-(OMe)2 was 2.5 h. The values for the MICHAELIS constants obtained for Z-D,L-Asp-(OMe)2 (Km = 0.576 mM) and PhAc-D,L-Asp-(OMe)2 (Km = 0.300 mM) showed a high affinity of the enzyme to the substrates. For Boc-D,L-Asp-(OMe)2 the affinity of the enzyme is considerable lower (Km = 14.07 mM). The results of these investigations can be effectively used for the separation of N-protected derivatives of D,L-aspartic acid and with a high probability also for other amino and racemic forms.  相似文献   

14.
Bovine descarboxyprothrombin and descarboxyfragment-1 can be used as substrates for rat and bovine vitamin K-dependent carboxylase. In both enzyme systems, however, these substrates have a high Km (0.3–0.4 mM). A better substrate (Km = 0.001–0.003 mM) was prepared from bovine descarboxyprothrombin by limited proteolysis with subtilisin Carlsberg. This substrate is called Fragment-Su and is composed of the amino acids 13–29 of descarboxyprothrombin.  相似文献   

15.
λ-Glutamylcysteine synthetase activity (EC 6.3.2.2) was analysed in Sephacryl S-200 eluents of extracts from cell suspension cultures ofNicotiana tabacum L. cv. Samsun by determination of λ-glutamylcysteine as its monobromobimane derivative. The enzyme has a relative molecular mass (Mr) of 60000 and exhibits maximal activity at pH 8 (50% at pH 7.0 and pH9.0) and an absolute requirement for Mg2+. With 0.2mM Cd2+ or Zn2+, enzyme activity was reduced by 35% and 19%, respectively. Treatment with 5 mM dithioerythritol led to a heavy loss of activity and to dissociation into subunits (Mr 34000). Buthionine sulfoximine andl-methionine-sulfoximine, known as potent inhibitors of λ-glutamylcysteine synthetase from mammalian cells, were found to be effective inhibitors of the plant enzyme too. The apparent Km values forl-glutamate,l-cysteine, and α-aminobutyrate were, respectively, 10.4mM, 0.19 mM, and 6.36 mM. The enzyme was completely inhibited by glutathione (Ki=0.42 mM). The data indicate that the rate of glutathione synthesis in vivo may be influenced substantially by the concentration of cysteine and glutamate and may be further regulated by feedback inhibition of λ-glutamylcysteine synthetase by glutathione itself. λ-Glutamylcysteine synthetase is, like glutathione synthetase, localized in chloroplasts as well as in the cytoplasm. Chloroplasts fromPisum sativum L. isolated on a Percoll gradient contained about 72% of the λ-glutamylcysteine synthetase activity in leaf cells and 48% of the total glutathione synthetase activity. In chloroplasts ofSpinacia oleracea L. about 61% of the total λ-glutamylcysteine synthetase activity of the cells were found and 58% of the total glutathione synthetase activity. These results indicate that glutathione synthesis can take place in at least two compartments of the plant cell. Dedicated to Professor A. Prison on the occasion of his 80th birthday  相似文献   

16.
Methylglyoxal reductase was purified from Hansenula mrakii IFO 0895 to a homogenous state on polyacrylamide gel electrophoresis. The enzyme consisted of a single polypeptide chain with a molecular weight of 34,000. The enzyme was specific to methylglyoxal (Km = 1.92 mM) and NADPH (Km = 40.8 μM). The activity of the enzyme was inhibited by p-chloromercuribenzoate and HgCl2. NADP also inhibited the activity of the enzyme, and the Ki value was calculated to be 0.25 mM.  相似文献   

17.
Apple fruit used for beverage production has a polyphenol oxidase which does not hydroxylate tyrosine under any conditions but it hydroxylates p-coumaric acid in the presence of NADH, and phloridzin in the absence of cofactors. The apparent Kms for hydroxylation of phloridzin and p-coumaric acid are 1.5 and 4 mM, respectively. However, subsequent oxidation of 3-hydroxyphloridzin or caffeic acid has an apparent Km of 200 nM. The oxidation products of 3-hydroxyphloridzin are complex and a stable dimeric quinone is finally formed. The apparent Kms for oxidation of catechin, epicatechin, chlorogenic acid, l-Dopa and 4-methyl catechol are 4.7, 5.7, 6.0, 2.7 and 3.2 mM, respectively. The Km for oxygen was 4.3 % although there was marked substrate inhibition by oxygen above 30 %. Polyphenol oxidase was stable at pH 3.5–4.5 with an optimum of 4.5.  相似文献   

18.
Nitrosopyrrolidine (NO-PYR), an hepatocellular carcinogen, is rapidly metabolized to CO2 by hepatocytes freshly isolated from the livers of male Fischer rats. Using CO2 evolution as a measure of NO-PYR metabolism, we observed two kinetic constants; a high affinity component (Km = 0.11 mM), and a lower affinity component (Km = 3.2 mM). The high affinity component has similar kinetic constants to those observed for in vitro reactions with microsomes plus cytosol (Km = 0.36 mM). Therefore, it is probable that the microsomal reaction is the limiting factor in the metabolism of NO-PYR in hepatocytes. NO-PYR may be metabolized to CO2 through normal anaplerotic sequences. Some metabolites of NO-PYR which have been tentatively identified are γ-hydroxybutyrate, succinic semialdehyde, 3,4-dihydroxybutyric acid lactone, lactate, acetate, pyruvate, glyoxylate, γ-aminobutyrate and alamine. 2-Hydroxytetrahydrofuran (2-hydroxy-THF), a product of α-hydroxylation was detected at low levels in only one of four reactions. 3-Hydroxy-NO-PYR is present but represents only a small percentage of the total metabolism and is probably of little significance in the overall catabolism of NO-PYR in hepatocytes.  相似文献   

19.
The kinetic features of glucose transport in human erythrocytes have been the subject of many studies, but no model is consistent with both the kinetic observations and the characteristics of the purified transporter. In order to reevaluate some of the kinetic features, initial rate measurements were performed at 0°C. The following kinetic parameters were obtained for fresh blood: zero-trans efflux Km = 3.4 mM, Vmax = 5.5 mM/min; infinite-trans efflux Km = 8.7 mM, Vmax = 28 mM/min. For outdated blood, somewhat different parameters were obtained: zero-trans efflux Km = 2.7 mM, Vmax = 2.4 mM/min; infinite-trans efflux Km = 19 mM, Vmax = 23 mM/min. The Km values for fresh blood differ from the previously reported values of 16 mM and 3.4 mM for zero-trans and infinite-trans efflux, respectively (Baker, G.F. and Naftalin, R.J. (1979) Biochim. Biophys. Acta 550, 474–484). The use of 50 mM galactose rather than 100 mM glucose as the infinite-trans sugar produced no change in the infinite-trans efflux Km values but somewhat lower Vmax values. Simulations indicate that initial rates were closely approximated by the experimental conditions. The observed time courses of efflux are inconsistent with a model involving rate-limiting dissociation of glucose from hemoglobin (Naftalin, R.J., Smith, P.M. and Roselaar, S.E. (1985) Biochim. Biophys. Acta 820, 235–249). The results presented here support the adequacy of the carrier model to account for the kinetics.  相似文献   

20.
The transport of thymidine and hypoxanthine was investigated in mutant Chinese hamster lung fibroblasts deficient in both thymidine kinase and hypoxanthine-guanine phosphoribosyltransferase. Kinetic data from rapid uptake experiments (0.5–4.5 s) indicate that thymidine is transported by a monophasic saturable system (Km = 0.29 mM, V = 6.7 nmol/min · mg) which is competitively inhibited by hypoxanthine (Ki = 3.3 mM). The cells displayed a single transport system for hypoxanthine (Km = 2.0 mM, V = 8.9 nmol/min · mg) that is inhibited competitively by thymidine (Ki = 0.43 mM). Both hypoxanthine and thymidine entry were noncompetively inhibited by nitrobenzylthioinosine, but thymidine transport was more sensitive. A kinetic model in which hypoxanthine and thymidine share a common transporter can account for the competitive inhibition and the observation that the inhibition constants are similar to the Michaelis constants.  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