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 regulation of homoserine dehydrogenase (HSD) activity (EC 1.1.1.3) by L-threonine, L-cysteine and K+ was examined using extracts of organs of soybean seedlings harvested 3, 6, 11, and 19 days after germination. K+ stimulated HSD activity from each source at least 2-fold. HSD activity was completely inhibited by 10 mM L-cysteine while 10 mM D-cysteine was not inhibitory. A progressive decrease in sensitivity of NAD-dependent HSD to inhibition by 10 mM L-threonine occurred in all organs except the leaf during the sampling period. This progressive decrease in sensitivity of the HSD to threonine inhibition was detected only when K+ was present in the assay mixtures. Four major molecular forms, including one rapidly migrating form (form I) and three more slowly migrating forms (forms II, III, IV) of HSD, were identified in extracts of soybean organs by polyacrylamide electrophoresis. Chromatographic and electrophoretic data indicate that form I, which was not inhibited by threonine or stimulated by K+, was of lower MW than forms II, III and IV which were of similar MW. These latter 3 forms were inhibited by threonine and stimulated by K+. During soybean seedling development form II increased in amount and forms I and IV decreased in amount. This alteration in the amounts of the forms of HSD occurred during the same period as the decrease in the amount of threonine inhibition. Since K+ stimulation of HSD decreased during soybean organ development and K+ enhanced threonine inhibition, this might account for the observed decrease in threonine inhibition.  相似文献   

2.
Phosphoglucoisomerase from cytosol of immature wheat endosperm was purified 650-fold by ammonium sulphate fractionation, isopropyl alcohol precipitation, DEAE-cellulose chromatography and gel filtration through Sepharose CL-6B. The enzyme, with a molecular weight of about 130,000, exhibited maximum activity at pH 8.1. It showed typical hyperbolic kinetics with both fructose 6-P and glucose 6-P withK m of 0.18 mM and 0.44mM respectively. On either side of the optimum pH, the enzyme had lower affinity for the substrates. Using glucose 6-P as the substrate, the equilibrium was reached at 27% fructose 6-P and 73% glucose 6-P with an equilibrium constant of 2.7. The ΔF calculated from the apparent equilibrium constant was +597 cal mol-1. The activation energy calculated from the Arrhenius plot was 5500 cal mol-1. The enzyme was completely inhibited by ribose 5-P, ribulose 5-P and 6-phosphogluconate, withK i values of 0.17, 0.25 and 0.14 mM respectively. The probable role of the enzyme in starch biosynthesis is discussed.  相似文献   

3.
Fructokinase has been purified from developing potato (Solanum tuberosum L.) tubers by a combination of hydrophobic interaction, affinity chromatography, and gel filtration. The protein has a native molecular mass of approximately 70 kD but is apparently a dimer. Ion-exchange chromatography and two-dimensional western blots resolved three major fructokinases, designated FK-I, FK-II, and FK-III in order of their elution from a Mono-Q column. Fructokinase activity proved labile when proteins were purified in the absence of fructose. Kinetically, FKs I, II, and III all have broad pH optima with peaks at about pH 8.5. The enzymes have a high specificity for fructose (Km values ranging from 0.041 to 0.128 mm), and can utilize a range of nucleoside triphosphates. Unlike FKs I and II, FK-III is not inhibited by fructose concentrations in excess of 1 mm. MgADP inhibited activity of the three FKs (between 68 and 75% inhibition at 1.0 mm), whereas fructose 6-P caused inhibition at concentrations of 10 mm. There were no regulatory effects observed with a range of other metabolites. K+ (10 mm) activated FK-I by 4-fold and FKs II and III by only about 50%.  相似文献   

4.
Both Pi-repressible acid phosphatases, IIb (mycelial) and IIc (extracellular), synthesized by Neurospora crassa and purified to apparent homogeneity by 7.5% PAGE, are monomers, are inhibited by 2 mm ZnCl2 and are non-specifically stimulated by salts. However, the IIc form is activated by p-nitrophenylphosphate (in a negative co-operativity effect with a K 0.5 of 2.5 mm) whereas form IIb shows Michaelis kinetics, with a K m of 0.5 mm. Thus, since both enzymatic forms may be expressed by the same gene (pho-3), it is possible that post-translational modifications lead to the excretion of an enzymatic form with altered Michaelis kinetics compared with the enzymatic form retained by the mycelium.  相似文献   

5.
Three Acid phosphatases (ACP) were isolated and characterized from the lysosomes of blood stream forms of Trypanosoma brucei by a combination of isopynic and differential centrifugation through Ficoll, organic solvent precipitation, ion exchange on DEAE cellulose 52 and size exclusion chromatography on Sephadex G-75 columns. The purified ACP emerged as three distinct peaks (ACP I, ACP II and ACP III) with high specific activities and they moved homogenously on 12% SDS-PAGE each as a single band with relative molecular weight of 36 kDa, 25 kDa and 45 kDa respectively. The purified enzymes were active at an optimum pH and temperature of 5.5 and 40 °C respectively. The enzyme activities appeared to be ACP because their activities were enhanced at low pH values and inhibited by the acid phosphatase inhibitor, sodium fluoride. ACP I and ACP II were sensitive to l-tartrate while ACP III was insensitive to l tartrate. The kinetic analysis of the purified enzyme (ACP I, ACP II and ACP III) determined using para-nitrophenylphosphate as substrate gave KM values of 0.2 mM, 0.15 mM and 0.5 mM. Monofunctional group sulfhydryl group inhibitors; HgCl2, and AgCl2 strongly inhibited the activity of ACP III and millimolar concentrations of dithiothreitol and iodoacetamide activated and inhibited the activity of the ACP III respectively, suggesting the involvement of thiol groups at the active site of the enzyme. Thus, differentiating it from ACP I and ACP II. The implication of these findings in relation to the pathology of trypanosomosis is discussed.  相似文献   

6.
Both of the starch phosphorylase fractions from Victory Freezer pea seeds, that can be separated by DEAE—cellulose chromatography and purified by Sepharose 4B-starch affinity chromatography, contain pyridoxal 5′-phosphate. The addition of further quantities of pyridoxal 5′-phosphate causes inactivation. Both enzymes showed similar bi-substrate kinetics with d-Glc-1-P and varying amounts of amylopectin and also with Pi and varying amounts of amylopectin. In the direction of glucan sythesis the Km for amylopectin with phosphorylase II was much higher than with phosphorylase I. However, the two enzymes differed in their behaviour on glucan degradation at varying concentrations of Pi. With phosphorylase II the Km for amylopectin was dependent on the concentration of Pi but that for phosphorylase I was constant. Phosphorylase II was strongly inhibited by ADPG in the direction of glucan degradation but only slightly in the direction of glucan synthesis by both ADPG and UDPG. Phosphorylase I was only slightly inhibited by ADPG in both directions and by UDPG in synthesis. UDPG inhibited both enzymes moderately in glucan degradation,  相似文献   

7.
Three forms of pyrophosphate fructose-6-phosphate 1-phosphotransferase (PFP) were purified from both green and red tomato (Lycopersicon esculentum) fruit: (a) a classical form (designated Q2) containing α- (66 kilodalton) and β- (60 kilodalton) subunits; (b) a form (Q1) containing a β-doublet subunit; and (c) a form (Q0) that appeared to contain a β-singlet subunit. Several lines of evidence suggested that the different forms occur under physiological conditions. Q2 was purified to apparent electrophoretic homogeneity; Q1 and Q0 were highly purified, but not to homogeneity. The distribution of the PFP forms from red (versus green) tomato was: Q2, 29% (90%); Q1, 47% (6%); and Q0, 24% (4%). The major difference distinguishing the red from the green tomato enzymes was the fructose-2,6-bisphosphate (Fru-2,6-P2)-induced change in Km for fructose-6-phosphate (Fru-6-P), the `green forms' showing markedly enhanced affinity on activation (Km decrease of 7-9-fold) and the `red forms' showing either little change (Q0, Q1) or a relatively small (2.5-fold) affinity increase (Q2). The results extend our earlier findings with carrot root to another tissue and indicate that forms of PFP showing low or no affinity increase for Fru 6-P on activation by Fru-2,6-P2 (here Q1 and Q0) are associated with sugar storage, whereas the classical form (Q2), which shows a pronounced affinity increase, is more important for starch storage.  相似文献   

8.
Two forms of p-nitrophenyl α-D-mannosidase and p-nitrophenyl α-D-galactosidase were purified from the protein bodies of mature Lupinus angustifolius seeds. A MW of 300 000 was calculated for both α-mannosidase A and B with Km = 1.92 and 2.70 mM and activation energies of 10.9 and 10.8 kcal/mol, respectively. α-Galactosidase I and II had MWs of 70800 and 17000 with Km = 0.282 and 0.556 mM and activation energies 17.7 and 11.5 kcal/mol, respectively. The enzymes had acid pH optima and were inhibited by various metal ions, carbohydrates and glycoproteins. They were able to release free sugar from several putative natural substrate oligosaccharides and the Lupinus storage glycoprotein, α-conglutin.  相似文献   

9.
Glycogen phosphorylase and synthase activities were detected in the sonic lysate of rumen ciliates of the genus Entodinium. The ciliate phosphorylase had the following properties. The pH optimum was narrow and centered at pH 5.9. The activity was maximum at 30°C; above 40°C a rapid inactivation occurred. The Km value for glucose-1-phosphate (G-1-P) and for glycogen was 15 mM and 0.069% (w/v), respectively. NaF and ethylenediamine tetraacetic acid had no stimulative effect on the enzyme activity, though adenosine 3′,5′-monophosphate and theophylline activated it. NaHSO3 inhibited the enzyme activity at a concentration of 1 mM. The inhibition of glucose was noncompetitive for G-1-P. Glycolytic intermediates and nucleotides had a minor effect on phosphorylase activity. Glycogen synthase existed in two forms, glucose-6-phosphate dependent and independent forms: the proportion of the latter form increased with the decrease of reserve polysaccharide levels in the ciliates. Correlations between glycolytic enzyme activities included phosphorylase and synthase activities and reserve polysaccharide contents in the ciliates were determined, and a possible regulatory mechanism of polysaccharide synthesis and degradation was discussed.  相似文献   

10.
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   

11.
Phosphoenolpyruvate carboxylase (PEPC) was purified from leaves of four species of Alternanthera differing in their photosynthetic carbon metabolism: Alternanthera sessilis (C3), A. pungens (C4), A. ficoides and A. tenella (C3-C4 intermediates or C3-C4). The activity and properties of PEPC were examined at limiting (0.05 mM) or saturating (10 mM) bicarbonate concentrations. The Vmax as well as Km values (for Mg2+ or PEP) of PEPC from A. ficoides and A. tenella (C3-C4 intermediates) were in between those of C3 (A. sessilis) and C4 species (A. pungens). Similarly, the sensitivity of PEPC to malate (an inhibitor) or G-6-P (an activator) of A. ficoides and A. tenella (C3-C4) was also of intermediate status between those of C3 and C4 species of A. sessilis and A. pungens, respectively. In all the four species, the maximal activity (Vmax), affinity for PEP (Km), and the sensitivity to malate (KI) or G-6-P (KA) of PEPC were higher at 10 mM bicarbonate than at 0.05 mM bicarbonate. Again, the sensitivity to bicarbonate of PEPC from C3-C4 intermediates was in between those of C3- and C4-species. Thus the characteristics of PEPC of C3-C4 intermediate species of Alternanthera are intermediate between C3- and C4-type, in both their kinetic and regulatory properties. Bicarbonate could be an important modulator of PEPC, particularly in C4 plants. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

12.
The carbonic anhydrases (CAs, EC 4.2.1.1) represent a superfamily of widespread enzymes, which catalyze a crucial biochemical reaction, the reversible hydration of carbon dioxide to bicarbonate and protons. Human CA isoenzymes I and II (hCA I and hCA II) are ubiquitous cytosolic isoforms. In this study, a series of hydroperoxides, alcohols, and acetates were tested for the inhibition of the cytosolic hCA I and II isoenzymes. These compounds inhibited both hCA isozymes in the low nanomolar ranges. These compounds were good hCA I inhibitors (Kis in the range of 24.93–97.99?nM) and hCA II inhibitors (Kis in the range of 26.04–68.56?nM) compared to acetazolamide as CA inhibitor (Ki: 34.50?nM for hCA I and Ki: 28.93?nM for hCA II).  相似文献   

13.
The fungus Humicola lutea 120-5 cultivated in casein-containing media, in the presence or absence of inorganic phosphate (Pi), excretes three different molecular forms of acid phosphatase (with Mr values of approximately 140, 70 and 35 kDa). The enzyme forms were isolated and purified 30–100-fold by a procedure involving two steps of ion-exchange chromatography and Sephadex G-200 gel chromatography. It was found that the fungus excretes only one of the phosphatases with the highest Mr (140 kDa) during growth on medium with inorganic nitrogen source (NaNO3). This form (designed AcPh I) was assumed to be a constitutive, since it showed resistance to high Pi-concentrations (10 mM) and its biosynthesis was not affected by the type of nitrogen source (casein or NaNO3). The other two forms (AcPh II-70 and AcPh III-35 kDa) were competitively inhibited by Pi (K i = 0.5 and 0.2 mM, respectively) and were induced by casein. The K m values of AcPh I and AcPh II were estimated as 1.3 mM, while AcPh III showed a higher affinity for p-nitrophenylphosphate (pNPP) with K m of 0.5 mM. The AcPh I–III fractions demonstrated a pH optimum in the range of 4.5–4.8 and an optimal temperature of 55 °C using pNPP as a substrate. This revised version was published online in November 2006 with corrections to the Cover Date.  相似文献   

14.
Using hydrophobic chromatography, Chlorella ferredoxin was separated into three components (Fd I, Fd II and Fd III) in ratios of approximately 3:13:1. The three components differed in isoelectric point, peptide mapping, amino acid composition and N-terminal sequence. Fd II and Fd III were found to support fairly high rates of cytochrome c reduction by spinach FNR, while Fd I could not support this reaction at all. The highest value of the specificity constant (kcat/Km for NiR was demonstrated for Fd II-dependent activity; however, the lowest value of kcat/Km for NiR was obtained using Fd II.  相似文献   

15.
In the presence of porcine submaxillary N-acetylgalactosaminyltransferase and uridine diphospho-N-acetyl-D-galactosamine, approx. 1.2–1.5 mol of N-acetylgalactosamine were transfered per mol of myelin basic protein. Tritium-labelled N-acetylgalactosamine-labelled basic protein was digested with trypsin and the peptides were separated by HPLC and the radioactivity measured. Most of the radioactivity was associated with three peptide peaks (I, II and III) containing 17, 69 and 6% of the total radioactivity, respectively. The remaining radioactivity was distributed amongst several peptides, each containing less than 2.5% of the total radioactivity. Glycosylation of the basic proteins isolated from human, bovine and guine pig myelins showed that they were all equally good acceptors. In spite of differences in the peptide profiles of the basic proteins from different species, the distribution of radioactivity between the three peptide peaks was similar for all the species studies. The transfer of N-acetylgalactosamine to peptide II was much faster than to peptides I and III. The apparent Km values of the three peptides were within a narrow range of 0.52–0.63 mM, whereas the Vmax values were considerably different. The glycosylated peptide peaks (I, II and III) were separated by electrophoresis, the radioactivity measured, and amino acid compositions determined after hydrolysis. The major radioactive peptides of the human basic protein were identified with tryptic peptides containing the following sequences:
  相似文献   

16.
Glycine enhanced the sensitivity of maize phosphenolpyruvate carboxylase to the activator glucose 6-phosphate and reduced the sensitivity of the enzyme to the inhibitors malate and aspartate. The effects of glycine on the kinetic constants for these other effectors were greater than its effect on the Km for substrate, raising the Ki(malate) 11-fold and reducing Ka(glucose6-P) 7-fold, while reducing the Km(PEP) by 3-fold. Kinetically saturating levels of glycine and glucose 6-phosphate acted synergistically to raise Ki(malate) higher than that observed with either activator alone. Glycine and glucose 6-phosphate also synergistically reduced aspartate inhibition. Dual inhibitor analysis indicated that aspartate and malate bind in a mutually exclusive manner, and thus probably compete for the same inhibitor site. In contrast, the synergism between glycine and glucose 6-phosphate indicate that these activators bind at separate sites. Glycine also reduced the Km(Mg) by 3-fold but had no significant effect on the Km of bicarbonate.Abbreviation PEP phosphoenolpyruvate  相似文献   

17.
Two serine racemases (I and II) were isolated from Streptomyces garyphalus. Serine racemase I (molecular weight 93,000) was purified to a single band in an analytical electrofocusing system. Serine racemase II (molecular weight 73,000) was partially purified. Both enzymes used pyridoxal-5-phosphate as cofactor. Besides serine the enzymes utilized alanine as substrate but no other amino acid tested. The K m values of l-alanine and l-serine for enzyme I were 111 mM and 35 mM respectively. Enzyme I was not inhibited by d-cycloserine but by hydroxylamine. Both substances inhibited enzyme II. The serine racemases may be involved in the biosynthesis of d-cycloserine in S. garyphalus.  相似文献   

18.
Marine Micrococcus luteus K-3 constitutively produced two salt-tolerant glutaminases, designated glutaminase I and II. Glutaminase I was homogeneously purified about approximately, 1620-fold with a 4% yield, and was a dimer with a molecular weight of about 86,000. Glutaminase II was partially purified about 190-fold with a 0.04% yield. The molecular weight of glutaminase II was also 86,000. Maximum activity of glutaminase I was observed at pH 8.0, 50°C and 8–16% NaCl. The optimal pH and temperature of glutaminase II were 8.5 and 50°C. The activity of glutaminase II was not affected by the presence of 8 to 16% NaCl. The presence of 10% NaCl enhanced thermal stability of glutaminase I. Both enzymes catalyzed the hydrolysis of l-glutamine, but not its hydroxylaminolysis. The Km values for l-glutamine were 4.4 (glutaminase I) and 6.5 mM (glutaminase II). Neither of the glutaminases were activated by the addition of 2 mM phosphate or 2 mM sulfate. p-Chloromercuribenzoate (0.01 mM) significantly inhibited glutaminase I, but not glutaminase II. The conserved sequences LA**V and V**GGT*A were observed in the N-terminal amino acid sequences of glutaminase I, similar to that for other glutaminases.  相似文献   

19.
Carbapenem-hydrolyzing β-lactamase from Serratia marcescens FHSM4055 was purified 926-fold by means of carboxylmethyl Sephadex C-50, Sephacryl S-200, and Mono S column chromatography. The molecular weight was 30,000 by SDS-PAGE and the isoelectric point was 8.7. The enzyme activity was inhibited by EDTA, and restored by adding zinc (II) or manganese (II). It was inhibited by p-chloromercuribenzoate and iodine as well as the heavy metals, Hg (II), Fe (II), Fe (III), and Cu (II). These results indicate that the enzyme is a metallo-β-lactamase and that the SH-group of only one cysteine residue probably binds to the metal ion, thus contributing to the stability of the enzyme active center. The specific constant (kcat/Km) showed that the enzyme hydrolyzed various β-lactam antibiotics such as carbapenems, cephalosporins, moxalactam, cephamycins, and penicillins other than monobactams. Ampicillin and piperacillin with respective amino- and imino-groups, ceftazidime with a carboxypropyloxyimino-group, and cefclidin with a carbamoylquinuclidine-group were poor substrates among the β-lactam antibiotics other than the monobactams tested. The plots of the turnover number (kcat) against pH for the hydrolysis of cephaloridine gave an asymmetrical curve with the ‘tail’ on the acid side (pK1, 5.9; pK2, 9.0; pK3, 10.8), whereas those of kcat/Km gave a bell-shaped curve (pK1, 5.8; pK2, 9.8). Both results suggest that two ionic forms of an intermediate yield the same product at different rates and that the enzyme is stable under alkaline conditions. Since the N-terminal amino acid sequence of 27 residues determined was consistent with that of the metalloenzyme (Antimicrob. Agents Chemother., 1994, 38: 71-78), the above enzymatic characteristics seem to coincide.  相似文献   

20.
Formaldehyde dehydrogenase and formate dehydrogenase were purified 45- and 16-fold, respectively, from Hansenula polymorpha grown on methanol. Formaldehyde dehydrogenase was strictly dependent on NAD and glutathione for activity. The K mvalues of the enzyme were found to be 0.18 mM for glutathione, 0.21 mM for formaldehyde and 0.15 mM for NAD. The enzyme catalyzed the glutathine-dependent oxidation of formaldehyde to S-formylglutathione. The reaction was shown to be reversible: at pH 8.0 a K mof 1 mM for S-formylglutathione was estimated for the reduction of the thiol ester with NADH. The enzyme did not catalyze the reduction of formate with NADH. The NAD-dependent formate dehydrogenase of H. polymorpha showed a low affinity for formate (K mof 40 mM) but a relatively high affinity for S-formylglutathione (K mof 1.1 mM). The K mvalues of formate dehydrogenase in cell-free extracts of methanol-grown Candida boidinii and Pichia pinus for S-formylglutathione were also an order of magnitude lower than those for formate. It is concluded that S-formylglutathione rather than free formate is an intermediate in the oxidation of methanol by yeasts.  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