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.
The activity of pure Escherichia coli murein (peptidoglycan) amidase (N-acetylmuramoyl-l-alanine amidase, EC 3.5.1.28) was measured after preincubation with E. coli phosphatidylglycerol microdispersions in final concentration ranging over micro- and millimolarities. The enzyme activity was increased up to 160% of the control for phosphatidylglycerol concentrations increasing from 2 to 50 μM. After a plateau extending from 0.05 to 0.3 mM, higher phosphatidylglycerol concentrations inactivated the enzyme down to 15% of initial activity for concentrations of 2 mM. Positive kinetic cooperativity was observed for the activation as well as for the inactivation processes. Cardiolipin (or diphosphatidylglycerol) from the same origin and under same conditions had no significant effect. Molecular sieving experiments have shown that, when inactivated, the enzyme remained firmly bound to the phosphatidylglycerol vesicles, whereas the activated phosphatidylglycerol-enzyme complex was totally dissociable by dilution. Activated phosphatidylglycerol complexes were recovered by gel exclusion chromatography at equilibrium in 40 μM phosphatidylglycerol. Possible physiological meaning of the results is briefly discussed in the context of our work and that done previously by others.  相似文献   

2.
A purified preparation of N-acetylmuramoyl-L-alanine amidase (EC 3.5.1.28), a murein hydrolase from Escherichia coli, was found to lose its activity during incubation in the presence of bacterial phospholipid suspensions. Whether it was co-dispersed with the phospholipids or added to sonicated phospholipid suspension, the enzyme was inhibited (or inactivated) from the first minutes of incubation at 37 degree C. As phosphatidylglycerol/cardiolipin ratio of the phospholipid suspension as increased (all other things being equal), a further decrease of amidase activity was observed. The highest losses of activity were found after co-dispersion of the enzyme and the substrate together with the phospholipids, the resulting suspension being formed of larger multilayered vesicles, as revealed by electron microscopy. In these conditions, the effect on enzyme activity was only partially accounted for by the proportion of the enzyme that was entrapped in the vesicles. The entrapment capacity of the enzyme (using a 35S-labelled enzyme preparation) and of the substrate (3H-labelled) by the multilamellar phospholipidic vesicles did not significantly change as a function of their relative content of phosphatidylglycerol and cardiolipin. The possible physiological meaning of the results is discussed is connection with our previous data and with other works related to membranous phospholipid distribution and to septum formation control in bacteria.  相似文献   

3.
The purification and in vitro inactivation of AMP-deaminase from white muscle of carp Cyprinus carpio were conducted in the Fe2+/H2O2 and Fe2+/ascorbate oxidation systems. The enzyme activity decreases by 50% within 30 minutes of incubation in the presence of 100 microM of hydrogen peroxide and 5 microM of ferrous sulfate. Inactivation depended on incubation time and concentrations of FeSO4 and H2O2. In the system Fe2+/ascorbate the enzyme activity decreased by 50% at concentration of ascorbate 1 mM and 5 ferrous sulfate microM. Sodium nitrite did not affect the activity. S(0.5) and n(H) of both native and partially inactivated enzymes were virtually the same, while maximal activity of the inactivated enzyme was 2-3-fold lower than that of the native one.  相似文献   

4.
Depending on its concentration, phosphatidylglycerol, one of the three main Escherichia coli phospholipid species, is able to activate or inactivate the E. coli murein amidase (N-acetylmuramoyl-L-alanine amidase, EC 3.5.1.28) (Vanderwinkel, E. and De Vlieghere, M. (1985) Biochim. Biophys. Acta 838, 54-59). The mechanisms underlying the modulation of this enzyme activity were studied by analyzing the effects of cations, polycationic molecules, various surfactants and amphiphilic water-soluble compounds. K+, Mg2+ and polyamines were all able to prevent completely the enzyme inactivation produced by millimolar order concentration of phosphatidylglycerol. The efficiencies of the ionic species tested were in the order K+ less than Mg2+ = putrescine less than spermidine less than spermine. The kinetics of the counteraction processes were all sigmoidal. By contrast, the activation of the murein amidase produced by phosphatidylglycerol in micromolar concentration appeared to be insensitive to the ionic strength of the medium. Surfactants and amphiphilic molecules differing in their polar head and hydrophobic tail were found to activate the enzyme at various degrees for concentrations below their critical micellar concentration. The non-ionic surfactants were the most potent activators and remarkably mimicked the phosphatidylglycerol activation. The enzyme activation process appeared to require only a hydrophobic solvation shell around the protein. All kinetic data supported our previous interpretation of the phosphatidylglycerol-enzyme interactions in terms of multisite non-allosteric theory.  相似文献   

5.
An enzyme was identified in human serum which unlike lysozyme cleaved the amide bond between N-acetyl-muramic acid and l-alanine of the peptide side chain of the rigid layer (murein) of Escherichia coli. The N-acetylmuramyl-l-alanine amidase released all of the peptide side chains including those to which the lipoprotein is bound. A portion of the peptide side chains of the Micrococcus lysodeikticus murein was also hydrolysed from the polysaccharide chains. E. coli, M. lysodeikticus, Bacillus subtilis and Staphylococcus aureus were not killed by the amidase. Treatment of E. coli with EDTA or osmotic shock rendered the cells sensitive to the amidase and they were killed. Possible biological functions of the amidase are discussed.The enzyme was separated from lysozyme in human serum. Gel permeation chromatography indicated a molecular weight of the active enzyme of 82,000 while gel electrophoresis in the presence of sodium dodecyl sulfate revealed a molecular weight of 75,000. Thus, the enzyme probably consists of a single polypeptide chain. Incubation with neuraminidase rendered the amidase more basic suggesting the release of sialic acid residues. The modified glycoprotein disclosed an increased activity to murein. Enzyme activity was inhibited by p-chloromercuribenzene sulfonate and ethyleneglycol-bis(2-aminomethyl) tetraacetate (EGTA) at 1 and 0.2 mM concentration, respectively, whereas EDTA up to 5 mM was without effect. The amidase was also inactivated by agents that reduce disulfide bridges.  相似文献   

6.
A method based on Fourier transform infrared spectroscopy (FT-IR) has been developed for assaying the Pseudomonas aeruginosa native amidase (E.C. 3.5.1.4), overproduced in an Escherichia coli strain. The kinetic of acetamide hydrolysis by the enzyme, in aqueous media, was monitored by measuring the intensity of the acetamide amide I band maximum at 1635 cm(-1) as a function of time. A value of 0.5mM(-1) cm(-1) was obtained for the extinction coefficient (epsilon) of acetamide at this frequency. The rate of the hydrolysis was found to be linear with the concentration of the enzyme up to 90 microM. The Michaelis-Menten kinetics parameters V and K(m) were determined as 30.7 U/mg and 4mM, respectively. These results were similar to those obtained using high-performance liquid chromatography analysis of the same hydrolytic reaction catalyzed by amidase either in water or in buffer. This suggests that the precision of the FT-IR method is suitable for the kinetic studies of amidase with the additional advantage of being able to perform a real-time measurement of the enzymatic activity.  相似文献   

7.
The ability of nine phospholipids to alter the activity of low-Km cyclic AMP phosphodiesterase was examined in microsomal fractions of rat adipocytes. The enzyme was activated by phosphatidylserine (21% at 300 microM) and phosphatidylglycerol (36% at 300 microM). The activation was concentration dependent over the range 1-1000 microM. Six other phospholipids were without effect. Phosphatidylinositol 4-phosphate inhibited the activity of the enzyme over the same range of concentrations (26% at 300 microM). Phosphatidylserine also activated a partially purified preparation of the enzyme, whereas phosphatidylinositol 4-phosphate was ineffective. The mechanism of the activation of the enzyme by phosphatidylserine and phosphatidylglycerol involved an increase in the apparent Vmax of the enzyme, while the inhibition by phosphatidylinositol 4-phosphate was associated with an increase in the Km of the enzyme for substrate. The phospholipid modulators of low-Km cyclic AMP phosphodiesterase activity did not alter the activity of high-Km cyclic AMP phosphodiesterase. The ability of phospholipids to alter the activity of low-Km cyclic AMP phosphodiesterase in native membranes suggests a possible role for phospholipids in metabolic regulation.  相似文献   

8.
Expression and purification of a recombinant enantioselective amidase   总被引:2,自引:0,他引:2  
Microbacterium sp. AJ115 metabolises a wide range of nitriles using the two-step nitrile hydratase/amidase pathway. In this study, the amidase gene of Microbacterium sp. AJ115 has been inserted into the pCal-n-EK expression vector and expressed in Escherichia coli BL21(DE3)pLysS. The expressed protein is active in E. coli and expression of the amidase gene allows E. coli to grow on acetamide as sole carbon and/or nitrogen source. Expression of active amidase in E. coli was temperature dependent with high activity found when cultures were grown between 20 and 30 degrees C but no activity at 37 degrees C. On induction, the amidase represents 28% of the total soluble protein in E. coli. The expressed amidase has been purified in a single step from the crude lysate using the calmodulin-binding peptide (CBP) affinity tag. The V(max) and K(m) of the purified enzyme with acetamide (50 mM) were 4.4 micromol/min/mg protein and 4.5mM, respectively. The temperature optimum was found to be 50 degrees C. Purified enzyme demonstrated enantioselectivity with the ability to preferentially act on the S enantiomer of racemic (R,S)-2-phenylpropionamide. S-2-phenylpropionic acid is produced with an enantiomeric excess of >82% at 50% conversion of the parent amide.  相似文献   

9.
E Wanker  A Huber    H Schwab 《Applied microbiology》1995,61(5):1953-1958
The enzyme levanase encoded by the sacC gene from Bacillus subtilis was overexpressed in Escherichia coli with the strong, inducible tac promoter. The enzyme was purified from crude E. coli cell lysates by salting out with ammonium sulfate and chromatography on DEAE-Sepharose CL-6B, S-Sepharose, and MonoQ-Sepharose. The purified protein had an apparent molecular mass of 75,000 Da in sodium dodecyl sulfate-polyacrylamide gel electrophoresis, which is in agreement with that expected from the nucleotide sequence. Levanase was active on levan, inulin, and sucrose with Km values of 1.2 microM, 6.8 mM, and 65 mM, respectively. The pH optimum of the enzyme acting on inulin was 5.5, and the temperature optimum was 55 degrees C. Levanase was rapidly inactivated at 60 degrees C, but activity could be retained for longer times by adding fructose or glycerol. The enzyme activity was completely inactivated by Ag+ and Hg2+ ions, indicating that a sulfhydryl group is involved. A ratio of sucrase to inulinase activity of 1.2 was found for the purified enzyme with substrate concentrations of 50 mg/ml. The mechanism of enzyme action was investigated. No liberation of fructo-oligomers from inulin and levan could be observed by thin-layer chromatography and size exclusion chromatography-low-angle laser light scattering-interferometric differential refractive index techniques. This indicates that levanase is an exoenzyme acting by the single-chain mode.  相似文献   

10.
The effects of chlorpromazine on various properties of the F1-ATPases from bovine heart mitochondria (MF1), the plasma membranes of Escherichia coli (EF1), and plasma membranes of the thermophilic bacterium PS3 (TF1) have been examined. While chlorpromazine inhibited MF1 with an I0.5 of about 50 microM and EF1 with an I0.5 of about 150 microM at 23 degrees C, the ATPase activity of TF1 was stimulated by chlorpromazine concentrations up to 0.6 mM at this temperature. Maximal activation of about 20% was observed at 0.2 mM chlorpromazine at 23 degrees C. Chlorpromazine concentrations greater than 0.6 mM inhibited TF1 at 23 degrees C. At 37 degrees C the ATPase activity of TF1 was doubled in the presence of 0.5 mM chlorpromazine, the concentration at which maximal stimulation was observed at this temperature. Chlorpromazine inhibited the rate of inactivation of EF1 by dicyclohexylcarbodiimide (DCCD) at 23 degrees C and pH 6.5. Concentrations of chlorpromazine which inhibited the ATPase activity of TF1 at pH 7.0 accelerated the rate of inactivation of the enzyme by DCCD at pH 6.5, while lower concentrations of the phenothiazine, which stimulated the ATPase, had no effect on DCCD inactivation. Chlorpromazine concentrations up to 1.0 mM had no effect on the rate of inactivation of TF1 by DCCD at 37 degrees C and pH 6.5. Chlorpromazine at 0.5 mM accelerated the rate of inactivation of MF1 by 5'-p-fluorosulfonylbenzoyladenosine (FSBA), while it slowed the rate of inactivation of EF1 by FSBA. The inactivation of TF1 by FSBA in the absence of chlorpromazine was complex and was not included in this comparison. Chlorpromazine protected MF1 and EF1 against cold inactivation. Whereas 100 microM chlorpromazine afforded about 90% stabilization of MF1 at 4 degrees C, only about 30% stabilization of EF1 was observed under the same conditions in the presence of 400 microM chlorpromazine. Each of the ATPases was inactivated by the structural analog of chlorpromazine, quinacrine mustard. Whereas 5 mM ATP and 5 mM ADP protected MF1 and TF1 against inactivation by 0.5 mM quinacrine mustard, the rate of inactivation of EF1 by quinacrine mustard was accelerated fourfold by 5 mM ATP and slightly accelerated by 5 mM ADP.  相似文献   

11.
The GlcNAc-1-P transferase was solubilized from pig aorta microsomal fractions using 0.5% Nonidet P-40. The activity of the solubilized enzyme was stimulated by exogeneously added phospholipids in the order phosphatidylglycerol greater than phosphatidylinositol greater than phosphatidylserine. When the enzyme was stored in 20% glycerol containing 20 micrograms of phosphatidylglycerol/mg of protein, more than 80% of the activity remained after storage for 6 days at 0-4 degrees C. On the other hand, in the absence of the stabilizers, the enzyme lost most of its activity within 24 h. The transferase was purified about 68-fold using ammonium sulfate and DEAE-cellulose fractionation. The DEAE-cellulose chromatography separated a heat-stable factor from the enzyme, which when added back to the partially purified enzyme stimulated about 5-fold. With this partially purified enzyme, the Km for UDP-GlcNAc was found to be 1 X 10(-7) M, and that for dolichyl-P about 1 X 10(-6) M. The stimulatory factor increased the Vmax for both UDP-GlcNAc and dolichyl-P 5-10-fold, but the Km values remained the same. The pH optimum for the enzyme was between 7.4 and 7.6, and either Mn2+ (1 mM) or Mg2+ (10 mM) was required for optimum activity. The GlcNAc-1-P transferase was also stimulated by the addition of GDP-mannose (or other purine sugar nucleotides) or dolichyl-phosphoryl-mannose to the incubation mixtures. These two compounds acted in different ways on the enzyme since their stimulatory effects were additive. The effect of GDP-mannose was found to be due to protection of the substrate, UDP-GlcNAc, from degradation, but the effect of dolichyl-P-mannose remains to be established. In addition, the stimulations shown by phosphatidylglycerol, GDP-mannose, and factor, or phosphatidylglycerol, dolichyl-P-mannose, and factor, were all additive, indicating that they were acting at different sites on the enzyme. The transferase was quite sensitive to the action of sulfhydryl reagents such as N-ethylmaleimide or p-chloromercuribenzene sulfonate, and was rapidly inactivated in their presence. The enzyme could be protected to the extent of about 50% when all of the substrates (UDP-GlcNAc, dolichyl-P, Mn2+) were added before the addition of the sulfhydryl reagents.  相似文献   

12.
The Swatek's method was further simplified for the assay of penicillin amidase activity. The absorbance of colour obtained during determination of 6-aminopenicillanic acid was dependent on concentration of 4-dimethylaminobenzaldehyde and on temperature. Antiodies induced in rabbits with one molecular form of penicillin amidase from E. coli PCM 271 (PA-1 or PA-2) did not cross-react with the other amidase form. No differences in substrate specificity on inactivation with SDS and in alkaline medium between the two amidase forms were observed. Concentrated urea inactivated PA-2 irreversibly and PA-1 reversibly. N-Bromosuccinimide inactivated almost completely only PA-1. Two E. coli PCM 271 strain variants were separated by microbial selection. Each of them produced only one amidase form. Also two amidase forms were found in cells of E. coli ATCC 11105, whereas E. coli ATCC 9636 and ATCC 9637 synthesize only PA-1.  相似文献   

13.
Partial purification and in vitro inactivation of glucose-6-phosphate dehydrogenase from the yeast Saccharomyces cerevisiae in the Fe2+/H2O2 oxidation system were conducted. At the protein concentration 1.5 mg/ml, the enzyme lost 50% of activity within 5 minutes of incubation in presence of 2 mM hydrogen peroxide and 3 mM ferrous sulphate. The inactivation extent depended on time and concentrations of FeSO4 and H2O2. EDTA, ADP and ATP at concentration 0.5 mM enhanced inactivation. At the same time, the presence of 0.5 mM NADPH, 1 mM glucose-6-phosphate, 10 mM mannitol, 30 mM dimethylsulphoxide or 20 mM urea diminished this process. In comparison with native enzyme, index S(0,5) of the partially inactivated enzyme for glucose-6-phosphate was 2.1-fold higher, but for NADP it was 1,6-fold lower. Maximal activity of the partially inactivated enzyme was 3-5-fold lower than that of native one.  相似文献   

14.
An amidase acting on (R,S)-piperazine-2-tert-butylcarboxamide was purified from Pseudomonas azotoformans IAM 1603 and characterized. The enzyme acted S-stereoselectively on (R,S)-piperazine-2-tert-butylcarboxamide to yield (S)-piperazine-2-carboxylic acid. N-terminal and internal amino acid sequences of the enzyme were determined. The gene encoding the S-stereoselective piperazine-2-tert-butylcarboxamide amidase was cloned from the chromosomal DNA of the strain and sequenced. Analysis of 2.1 kb of genomic DNA revealed the presence of two ORFs, one of which (laaA) encodes the amidase. This enzyme, LaaA is composed of 310 amino acid residues (molecular mass 34 514 Da), and the deduced amino acid sequence exhibits significant similarity to hypothetical and functionally characterized proline iminopeptidases from several bacteria. The laaA gene modified in the nucleotide sequence upstream from its start codon was overexpressed in Escherichia coli. The activity of the recombinant LaaA enzyme in cell-free extracts of E. coli was 13.1 units.mg(-1) with l-prolinamide as substrate. This enzyme was purified to electrophoretic homogeneity by ammonium sulfate fractionation and two column chromatography steps. On gel-filtration chromatography, the enzyme appeared to be a monomer with a molecular mass of 32 kDa. It had maximal activity at 45 degrees C and pH 9.0, and was completely inactivated in the presence of phenylhydrazine, Zn2+, Ag+, Cd2+ or Hg2+. LaaA had hydrolyzing activity toward L-amino acid amides such as L-prolinamide, L-proline-p-nitroanilide, L-alaninamide and L-methioninamide, but did not act on the peptide substrates for the proline iminopeptidases despite their sequence similarity to LaaA. The enzyme also acted S-stereoselectively on (R,S)-piperidine-2-carboxamide, (R,S)-piperazine-2-carboxamide and (R,S)-piperazine-2-tert-butylcarboxamide. Based on its specificity towards L-amino acid amides, the enzyme was named L-amino acid amidase. E. coli transformants overexpressing the laaA gene could be used for the S-stereoselective hydrolysis of (R,S)-piperazine-2-tert-butylcarboxamide.  相似文献   

15.
Various experiments were carried out in an attempt to determine the possible physiological function of the N-acetylmuramoyl-L-alanine amidase purified from Escherichia coli K12 on the basis of its activity on N-acetylmuramoyl-L-alanyl-D-gamma-glutamyl-meso-diaminopimelic acid [MurNAc-LAla-DGlu(msA2pm)]. A Km value of 0.04 mM was determined with this substrate. Specificity studies revealed that compounds with a MurNAc-LAla linkage are the most probable substrates of this enzyme in vivo. Purified amidase had no effect on purified peptidoglycan and only low levels (1-2.5%) of cleaved MurNAc-LAla linkages were detected in peptidoglycan isolated from normally growing cells. However, the action of the amidase in vivo on peptidoglycan was clearly detectable during autolysis. The amidase activity of cells treated by osmotic shock, ether or toluene, as well as that of mutants with altered outer membrane composition was investigated. Attempts to reveal a transfer reaction catalysed by amidase were unsuccessful. Furthermore, by its location and specificity, amidase was clearly not involved in the formation of UDP-MurNAc. The possibility that it might be functioning in vivo as a hydrolase degrading exogeneous peptidoglycan fragments in the periplasma was substantiated by the fact that MurNAc itself and MurNAc-peptides could sustain growth of E. coli as sole carbon and nitrogen sources. Finally, out of 200 thermosensitive mutants examined for altered amidase activity, only two strains had less than 50% of the normal level of activity, whereas ten strains were found to possess more than 50%. In fact, two of the overproducers encountered presented a 4-5-fold increase in activity.  相似文献   

16.
We describe an assay to measure the extent of enzymatic unwinding of DNA by a DNA helicase. This assay takes advantage of the quenching of the intrinsic protein fluorescence of Escherichia coli SSB protein upon binding to ssDNA and is used to characterize the DNA unwinding activity of recBCD enzyme. Unwinding in this assay is dependent on the presence of recBCD enzyme and linear dsDNA, is consistent with the known properties of recBCD enzyme, and closely parallels other methods for measuring recBCD enzyme helicase activity. The effects of varying temperature, substrate concentrations, enzyme concentration, and mono- and divalent salt concentrations on the helicase activity of recBCD enzyme were characterized. The apparent Km values for recBCD enzyme helicase activity on linear M13 dsDNA molecules at 25 degrees C are 0.6 nM dsDNA molecules and 130 microM ATP, respectively. The apparent turnover number for unwinding is approximately 15 microM base pairs s-1 (microM recBCD enzyme)-1. When this rate is corrected for the observed stoichiometry of recBCD enzyme binding to dsDNA, kcat for helicase activity corresponds to an unwinding rate of approximately 250 base pairs of DNA s-1 (functional recBCD complex)-1 at 25 degrees C. At 37 degrees C, the apparent Km value for dsDNA molecules was the same as that at 25 degrees C, but the apparent turnover number became 56 microM base pairs s-1 (microM recBCD enzyme)-1 [or 930 base pairs s-1 (functional recBCD complex)-1 when corrected for observed stoichiometry]. With increasing NaCl concentration, kcat peaks at 100 mM, and the apparent Km value for dsDNA increases by 3-fold at 200 mM NaCl. In the presence of 5 mM calcium acetate, the apparent Km value is increased by 3-fold, and kcat decreased by 20-30%. We have also shown that recBCD enzyme molecules are able to catalytically unwind additional dsDNA substrates subsequent to initiation, unwinding, and dissociation from a previous dsDNA molecule.  相似文献   

17.
Acetate kinase (ATP:acetate phosphotransferase, EC 2.7.2.1) from Escherichia coli exhibited a time-dependent loss of activity when incubated with N-ethylmaleimide at micromolar concentrations. However, prolonged incubation did not eliminate all catalytic activity and generally about 15% of its initial activity remained. When incubated with 7.2 microM N-ethylmaleimide, acetate kinase was inactivated with a rate constant of 0.063 min-1. Adenine nucleotides, ATP, ADP and AMP, protected the enzyme against such inactivation, but acetate up to 3.0 M and in the presence of 0.2 M MgCl2 and acetyl phosphate at 24 mM did not interfere with the rate of inactivation. While both acetate and acetyl phosphate did not affect the protection rendered by AMP, the presence of acetyl phosphate altered ADP protection. However, both substrates prevented ATP from protecting the enzyme. These data suggest that the binding sites for acetate and acetyl phosphate are different from that of the adenosine binding domain, but are in close vicinity to the phosphoryl binding regions of the nucleotides.  相似文献   

18.
Pep 5 and nisin are cationic peptide antibiotics which in addition to their membrane-disruptive action induce autolysis in staphylococci. To investigate the mechanism of lysis induction, the influence of the peptides on the activity of the N-acetylmuramoyl-L-alanine amidase of Staphylococcus simulans 22 was studied. In experiments with isolated cell walls at low ionic strength, the amidase activity was stimulated by the addition of Pep 5 and nisin, as well as by polylysine, streptomycin, and mono- and divalent cations. The concentrations necessary for activation depended on the nature of the cation and ranged from 5 microM for poly-L-lysine (n = 17) to 150 mM for Na+ at a cell wall concentration of 100 micrograms of cell walls per ml. No effect was observed if the cell walls were devoid of polyanionic constituents. Kinetic data suggested that the amidase bound to the teichoic and teichuronic acids of the cell wall and was thereby inhibited. Cationic molecules reversed this inhibition, most likely by displacing the enzyme from the polyanions. If the concentrations of the larger peptides were high in relation to cell wall concentration, the activation turned into inhibition, presumably by interfering with the access of the enzyme to its substrate. These experiments demonstrate that the activity of the amidase is modulated by basic peptides in vitro and help to explain how Pep 5 and nisin may cause lysis of treated cells.  相似文献   

19.
NADP-dependent isocitrate dehydrogenase activity has been screened in several cyanobacteria grown on different nitrogen sources; in all the strains tested isocitrate dehydrogenase activity levels were similar in cells grown either on ammonium or nitrate. The enzyme from the unicellular cyanobacterium Synechocystis sp. PCC 6803 has been purified to electrophoretic homogeneity by a procedure that includes Reactive-Red-120-agarose affinity chromatography and phenyl-Sepharose chromatography as main steps. The enzyme was purified about 600-fold, with a yield of 38% and a specific activity of 15.7 U/mg protein. The native enzyme (108 kDa) is composed of two identical subunits with an apparent molecular mass of 57 kDa. Synechocystis isocitrate dehydrogenase was absolutely specific for NADP as electron acceptor. Apparent Km values were 125, 59 and 12 microM for Mg2+, D,L-isocitrate and NADP, respectively, using Mg2+ as divalent cation and 4, 5.7 and 6 microM for Mn2+, D,L-isocitrate and NADP, respectively, using Mn2+ as a cofactor. The enzyme was inhibited non-competitively by ADP (Ki, 6.4 mM) and 2-oxoglutarate, (Ki, 6 mM) with respect to isocitrate and in a competitive manner by NADPH (Ki, 0.6 mM). The circular-dichroism spectrum showed a protein with a secondary structure consisting of about 30% alpha-helix and 36% beta-pleated sheet. The enzyme is an acidic protein with an isoelectric point of 4.4 and analysis of the NH2-terminal sequence revealed 45% identity with the same region of Escherichia coli isocitrate dehydrogenase. The aforementioned data indicate that NADP isocitrate dehydrogenase from Synechocystis resembles isocitrate dehydrogenase from prokaryotes and shows similar molecular and structural properties to the well-known E. coli enzyme.  相似文献   

20.
Spermine in micromolar concentrations decreased the basal activity of a guanosine 3',5'-monophosphate (cGMP) phosphodiesterase from bovine brain but had no effect in the presence of Ca2+ plus the calcium-dependent regulatory protein (CDR) which increased the activity of the enzyme 4- to 6-fold. Similar effects of spermine were observed on the enzyme at several stages of purification. Spermidine and putrescine were also inhibitory but higher concentrations were required. In the absence of Ca2+ and CDR, the enzyme exhibited two apparent Km values for cGMP (2.5 and 20 microM) which were unaltered by spermine. In the presence of Ca2+ and CDR (when spermine had no effect on activity), a single Km (3.5 microM) was observed. Enzyme purified by chromatography on CDR-Sepharose was rapidly inactivated during incubation at 30 degrees C in 5 mM potassium phosphate buffer (pH 7.0) with EDTA and ethylene glycol bis(beta-aminoethyl ether) N,N,N',N'-tetraacetic acid (EGTA). Spermine (20 microM) partially stabilized enzyme activity under these conditions, although it was somewhat less effective than 2 mM MgCl2. The inhibitory effects of spermine (or other polyamines) on basal phosphodiesterase activity, which can be overcome by Ca2+ and CDR, could be important in the regulation of cellular cyclic nucleotide content.  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