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.
在果糖1,6—二磷酸酯酶中果糖2,6—二磷酸可能与底物抑制的作用方式不同,因为蛇肌果糖1,6-二磷酸酯酶pH9.2的活性受到果糖2,6-二磷酸的抑制,而不受高浓度底物的影响。K+能增强果糖2,6—二磷酸对酶活性抑制,并能较大程度地解除过量底物的抑制。快反应流基修饰酶不再受较低浓度果糖2,6—二磷酸的抑制,但高浓度果糖2,6—二磷酸仍能抑制酶活性,其IC50增大40倍。修饰酶受底物抑制的阈值不变。为胰蛋白酶或枯草杆菌蛋白酶限制性酶解的果糖1,6—二磷酸酯酶受过量底物和果糖2,6—二磷酸抑制的行为也不相同。以上结果可能提示在蛇肌果糖1,6—二磷酸酯酸中存在既有别于AMP,又有别于过量底物的结合部位。  相似文献   

2.
Lactobacillus casei allosteric L-lactate dehydrogenase (L-LDH) absolutely requires fructose 1,6-bisphosphate [Fru(1,6)P2] for its catalytic activity under neutral conditions, but exhibits marked catalytic activity in the absence of Fru(1,6)P(2) under acidic conditions through the homotropic activation effect of substrate pyruvate. In this enzyme, a single amino acid replacement, i.e. that of His205 conserved in the Fru(1,6)P(2)-binding site of certain allosteric L-LDHs of lactic acid bacteria with Thr, did not induce a marked loss of the activation effect of Fru(1,6)P(2) or divalent metal ions, which are potent activators that improve the activation function of Fru(1,6)P(2) under neutral conditions. However, this replacement induced a great loss of the Fru(1,6)P(2)-independent activation effect of pyruvate or pyruvate analogs under acidic conditions, consequently indicating an absolute Fru(1,6)P(2) requirement for the enzyme activity. The replacement also induced a significant reduction in the pH-dependent sensitivity of the enzyme to Fru(1,6)P(2), through a slight decrease and increase of the Fru(1,6)P(2) sensitivity under acidic and neutral conditions, respectively, indicating that His205 is also largely involved in the pH-dependent sensitivity of L.casei L-LDH to Fru(1,6)P(2). The role of His205 in the allosteric regulation of the enzyme is discussed on the basis of the known crystal structures of L-LDHs.  相似文献   

3.
In Lactobacillus plantarum non-allosteric L-lactate dehydrogenase (L-LDH), the highly conserved His188 residue, which is involved in the binding of an allosteric effector, fructose 1,6-bisphosphate [Fru(1,6)P2], in allosteric L-LDH is uniquely substituted by an Asp. The mutant L. plantarum L-LDH, in which Asp188 is replaced by a His, showed essentially the same Fru(1,6)P2-independent catalytic activity as the wild-type enzyme, except that the Km and Vmax values were slightly decreased. However, the addition of Fru(1,6)P2 induced significant thermostabilization of the mutant enzyme, as in the case of many allosteric L-LDHs, while Fru(1,6)P2 showed no significant effect on the stability of the wild-type enzyme, indicating that only the single-point mutation, G-->C, sufficiently induces the Fru(1,6)P2-binding ability of L. plantarum L-LDH. The mutant enzyme showed higher thermostability than the wild-type enzyme in the presence of Fru(1,6)P2. In the absence of Fru(1,6)P2, on the other hand, the mutant enzyme was more labile below 65 degrees C but more stable above 70 degrees C.  相似文献   

4.
5.
T H Murcott  H Gutfreund    H Muirhead 《The EMBO journal》1992,11(11):3811-3814
The cooperative binding of the allosteric activator fructose-1,6-bisphosphate [Fru(1,6)P2] to yeast pyruvate kinase was investigated by equilibrium dialysis and fluorescence quench titration. The results show that yeast pyruvate kinase binds four molecules of Fru(1,6)P2 per tetramer and the observed fluorescence quench follows the binding of the ligand and not the cooperative T to R state transition. Additionally it is shown that the binding of Fru(1,6)P2 to yeast pyruvate kinase is compatible with the model of cooperativity that has been proposed and incorporates an intermediate state, R', with properties between those of the T and R states.  相似文献   

6.
Tetrameric 20 beta-hydroxysteroid dehydrogenase (17,20 beta,21-trihydroxysteroid:NAD+ oxidoreductase, EC 1.1.1.53) from Streptomyces hydrogenans was reactivated after inactivation, dissociation and denaturation with urea. The effect of several factors such as NAD+, NADH, substrate, sulphydryl reducing agents, extraneous proteins, pH and enzyme concentration on reactivation was investigated. The coenzymes were found to be essential for obtaining a high reactivation yield (about 90%), since in their absence the reactivation was less than 10%. NADH was effective at lower concentrations than NAD+. The reactivated enzyme, after the removal of inactive aggregates, showed physical and catalytic properties coincident with those of the native enzyme. The mechanism by which NADH affects the reconstitution of 20 beta-hydroxysteroid dehydrogenase was investigated using both soluble enzyme and enzyme immobilized on Sepharose 4B. The immobilization demonstrates that isolated subunits are inactive and incapable of binding NADH and suggests that subunit association to the tetramer is essential for enzymatic activity. NADH appears to act, after subunit assembly has taken place, by stabilizing tetramers and preventing their aggregation with monomers that would give rise to inactive polymers.  相似文献   

7.
The effect of the lyotropic series of anions on the stability and renaturation of tetrameric 20 beta-hydroxysteroid dehydrogenase (17,20 beta,21-trihydroxysteroid:NAD+ oxidoreductase, EC 1.1.1.53) was investigated. The variations in enzymatic activity were correlated with the changes in protein fluorescence, circular dichroism, reactivity of histidine residues and molecular weight. High concentrations of salting-out anions (phosphate, citrate, sulphate) were found to stabilize the enzyme markedly and increase the renaturation yield of the urea-denatured enzyme. Phosphate, for instance, induced the highest stabilization at about 1.2 M and the maximum reactivation (66%) at 0.5 M. At low anion concentration (0.01 M), the reactivation was only 7%. The renaturation property of salting-out anions seems to be due to their stabilizing effect on the end-product, i.e., the assembled tetramer. Salting-in anions (perchlorate, thiocyanate, iodide) inactivated the enzyme. At moderate anion concentrations (no greater than 0.25 M) the activation, which occurred slowly, without tetramer dissociation and with minor modifications of enzyme conformation, was fully reversed by concentrated phosphate or by saturating concentrations of NADH. In contrast, the inactivation induced by high anion concentrations (1-2 M) was rapid, irreversible and linked to considerable modifications of enzyme conformation.  相似文献   

8.
It has been investigated whether diurnal rhythms of sucrose-phosphate synthase (SPS) are involved in controlling the rate of photosynthetic sucrose synthesis. Extracts were prepared from spinach (Spinacia oleracea L.) and barley (Hordeum vulgare L.) leaves and assayed for enzyme activity. The activity of SPS increased in parallel with a rising rate of photosynthesis, and was increased by feeding mannose and decreased by supplying inorganic phosphate. In leaf material where sucrose had accumulated during the photoperiod or when sucrose was supplied exogenously, SPS activity decreased. During a diurnal rhythm, SPS activity increased after illumination, declined gradually during the light period, decreased further after darkening and then recovered gradually during the night. These changes did not involve an alteration of the maximal activity, but were caused by changes in the kinetic properties, revealed as a change in sensitivity to inhibition by inorganic phosphate. In experiments which modelled the response of SPS to changing metabolite concentrations, it was shown that these alterations of kinetic properties would strongly modify the activity of SPS in vivo. It is proposed that SPS can exist in kinetically distinct forms in vivo, and that the distribution between these forms can be rapidly altered. As the rate of photosynthesis increases there is an activation of SPS, which may be directly or indirectly linked to changes in the availability of Pi. This activation can be modified by factors related to the accumulation of sucrose. Under normal conditions there is a balance between these factors, and the leaf contains a mixture of the different forms of SPS.Abbreviations Chl chlorophyll - Frul,6bisP fructose-1,6-bisphosphate - Fru2,6bisP fructose-2,6-bisphosphate - Fru6P fructose-6-phosphate - Fru1,6bisPase fructose-1,6-bisphosphatase - Fru6P 2kinase fructose-6-phosphate, 2kinase - Fru2,6bisPase fructose-2,6-bisphosphatase - Glc6P glucose-6-phosphate - Pj inorganic phosphate - SPS sucrose-phosphate synthase - UDPGLc uridine 5-diphosphate glucose  相似文献   

9.
Homogeneous liver pyruvate kinase was reacted with different sulfhydryl reagents, which included o-iodosobenzoate, 5',5'-dithiobis(2-nitrobenzoic acid) and N-ethylmaleimide. Activity determinations of the treated enzyme made with and without Fru(1,6)P2 indicate that the protein contains two sulfhydryl groups per subunit important to its properties, one more accessible than the other. Fru(1,6)P2 added to mixtures prevented loss of activity obtained with o-iodosobenzoate and 5',5'-dithiobis(2-nitrobenzoic acid). It appears that Fru(1,6)P2 does not interfere with the reaction of the reagent with the sulfhydryl group, but prevents an ensuing conformational change, which leads to changes in the enzyme's properties.  相似文献   

10.
A cDNA encoding fructose(1,6)bisphosphatase was isolated from total human lung RNA. The cDNA contained an open reading frame encoding 337 amino acids. The determined nucleotide sequence of the lung cDNA was significantly different from muscle cDNA and slightly differed from human liver cDNA in a single mutation (Gly-336 for Ala-336) and a T for C substitution in position 648. The human lung fructose(1, 6)bisphosphatase [Fru(1,6)Pase] was isolated and its kinetic parameters were compared with liver and muscle isoenzymes. Values of kcat for the lung Fru(1,6)Pase were lower than for the liver and muscle enzyme. Like the liver isoenzyme, lung Fru(1,6)Pase is significantly less inhibited by AMP than the muscle enzyme. The values of I0.5 were 9.5, 9.8, and 0.3 microM for the liver, lung, and muscle enzyme, respectively. The lung enzyme was slightly more sensitive to fructose(2,6)bisphosphate [Fru(2,6)P2] inhibition than the liver enzyme. Ki was 75 microM for the lung and 96 microM for the liver enzyme. The synergistic effect of AMP and Fru(2,6)P2 on the lung and liver Fru(1,6)Pase was also observed. In the presence of AMP the corresponding values of Ki for Fru(2,6)P2 were 16 microM for the lung and 10 microM for the liver enzyme.  相似文献   

11.
The lactate dehydrogenase from Streptococcus faecalis is activated either by fructose 1,6-bisphosphate or by divalent cations such as Mn2+ or Co2+. With both types of activator, a lag is observed before attainment of the steady state rate of pyruvate reduction if the activator is added to the enzyme at the same time as the substrates. This lag can be largely abolished by preincubation of enzyme with activator before mixing with substrates. For fructose 1,6-bisphosphate (Fru(1,6)P2) as the activator, the rate constant for the lag phase showed a linear dependence on activator concentration but was independent of enzyme concentration. This suggests that binding of fructose 1,6-bisphosphate induces a conformational change in the enzyme which leads to increased activity, without association of enzyme subunits or dimers. With Co2+ as activator, the rate constant for the lag phase showed a hyperbolic dependence on Co2+ concentration and was also dependent on enzyme concentration. This suggests that activation by Co2+, in contrast to that by Fru(1,6)P2, involves association of enzyme dimers, followed by ligand binding.  相似文献   

12.
Purification and characterization of mitochondrial malate dehydrogenase [EC 1.1.1.37] from unfertilized eggs of the sea urchin, Anthocidaris crassispina, are described. The purification method consisted of dextran sulfate fractionation, Blue Dextran Sepharose chromatography, Phenyl-Sepharose hydrophobic chromatography and DEAE-cellulose chromatography. The enzyme was purified 771-fold with a 7% yield from the crude extract. The purified enzyme appeared homogeneous on polyacrylamide gel electrophoresis under both native and denatured conditions. After incubation at 45 degrees C for 50 min, the enzyme lost about 90% of its activity. In the presence of NADH, however, the enzyme was protected against the heat denaturation. The native enzyme had a molecular weight of about 65,000 and probably consisted of two identical subunits. In the reduction of oxaloacetate with NADH, a broad optimum pH ranging from 8.2 to 9.4 was found with 50 mM Tris-HCl and glycine-NaOH buffers. Sodium phosphate buffer apparently activated the enzyme. The apparent Km values for oxaloacetate and NADH were 19 microM and 30 microM, respectively. The optimum pH for malate oxidation with NAD+ was 10.2 in 50 mM NaHCO3-Na2CO3 buffer. The apparent Km values for malate and NAD+ were 7.0 mM and 0.6 mM, respectively. Zinc ion, sulfite ion, p-chloromercuriphenylsulfonate and adenine nucleotides strongly inhibited the enzyme.  相似文献   

13.
Blue Dextran has been coupled covalently to Sepharose-4B to purify the enzymatic complex NAD(P)H-nitrate reductase (EC 1.6.6.2) from the green alga Ankistrodesmus braunii by affinity chromatography. The optimum conditions for the accomplishment of the chromatographic process have been determined. The adsorption of nitrate reductase on Blue Dextran Sepharose is optimum when a phosphate buffer of low ionic strength and pH 6.5-7.0 is used. Once the enzyme has been bound to Blue Dextran Sepharose, it can be specifically eluted by addition of NADH and FAD to the washing buffer. However, none of the nucleotides added separately is able to promote the elution of the enzyme from the column. The elution can be also achieved, but not specifically, by increasing the ionic strength of the buffer with KCl. These results have made possible a procedure for the purification of A. braunii nitrate reductase which led to electrophoretic homogeneity, with an overall yield of 70% and a specific activity of 49 units/mg of protein.  相似文献   

14.
Activation of glycolysis by insulin in cultured adult rat hepatocytes is accompanied by an activation of phosphofructokinase 2 (PFK 2). PFK 2 activation might be caused by insulin-dependent changes of (a) metabolite levels, (b) basal and (c) Br8cAMP-stimulated cAMP-dependent protein kinase activity; this problem was investigated. 1. Cells cultured with 0.1 nM insulin for 48 h exhibited a low glycolytic rate and low fructose 2,6-bisphosphate [Fru(2,6)P2] levels. Addition of insulin increased Fru(2,6)P2 and Fru(1,6)P2 levels sequentially which points to PFK 2 as first target enzyme of insulin action. 2. Concentrations of Glc6P, Fru6P, phosphoenolpyruvate, glycerol 3-phosphate and citrate, which modulate PFK 2/fructose 2,6-bisphosphatase 2 activity, were not altered by insulin. 3. Activation of PFK 2 by insulin occurred without changes in the levels of total and protein-bound cAMP. Bound cAMP amounted to about 14% of total cAMP. 4. Insulin neither decreased the basal dissociation state of the cAMP-dependent protein kinase nor lowered the sensitivity of the kinase towards cAMP in cell extracts. 5. Addition of the phosphodiesterase-resistant Br8cAMP to the cultures increased cAMP levels 3-4-fold, elevated the protein kinase activity ratio from 0.14 to 0.6 and decreased the Fru(2,6)P2 level and the rate of glycolysis. When Br8cAMP and insulin were given together, insulin was capable of counteracting Br8cAMP in that it activated glycolysis and PFK 2 and elevated the Fru(2,6)P2 level; however, it did not decrease the elevated protein kinase activity ratio. It is concluded that insulin presumably does not activate PFK 2 through changes in cAMP and effector levels or through inhibition of cAMP-dependent protein kinase dissociation. The data support the hypothesis that insulin may act via activation of PFK 2 phosphatase.  相似文献   

15.
The aim of this work was to identify which aspects of photosynthetic metabolism respond most sensitively to leaf water deficit. Spinach (Spinacia oleracea L.) leaf discs were floated on sorbitol concentrations of increasing molarity and changes of the protoplast volume were estimated using [14C]sorbitol and 3H2O penetration. Detached leaves were also wilted until 10% of their fresh weight was lost. Photosynthesis was studied at very high external CO2 concentrations, to eliminate the effect of closing stomata. There was no large inhibition of CO2 fixation after wilting leaves, or until the external water deficit was greater than-1.2 MPa. However, partitioning changed markedly at these moderate water deficits: more sucrose and less starch was made. When an inhibition of CO2-saturated photosynthesis did appear at a water deficit of-2.0 MPa and above, measurements of chlorophyll-fluorescence quenching and metabolite levels showed the thylakoid reactions were not especially susceptible to short-term water stress. The inhibition was accompanied by a small increase of the triose phosphate: ribulose-1,5-bisphosphate ratio, showing regeneration of ribulose-1,5-bisphosphate was affected. However, there was also a general increase of the estimated concentrations of most metabolites, indicating that there is no specific site for the inhibition of photosynthesis. Increasing water deficit led to a large increase of fructose-2,6-bisphosphate. This is explained in terms of a simultaneous increase of fructose-6-phosphate and inorganic phosphate as the cell shrinks. The high fructose-2,6-bisphosphate led to the accumulation of triose phosphates, and the potential significance of this for protection against photoinhibition is discussed. There was an increase in the extractable activity of sucrose-phosphate synthase. This was only detected when the enzyme was assayed in conditions which distinguish between different kinetic forms which have previously been identified in spinach leaves. It is proposed that activation of sucrose-phosphate synthase is one of the first sites at which spinach leaves respond to a rising water deficit. This could be of importance for osmoregulation.Abbreviations Chl chlorophyll - Fru1,6bisP fructose-1,6-bisphosphate - Fru2,6bisP fructose-2,6-bisphosphate - Fru6P fructose-6-phosphate - Glc6P glucose-6-phosphate - PGA glycerate-3-phosphate - Pi inorgamic phosphate - Ru1,5bisP ribulose-1,5-bisphosphate - SPS sucrose-phosphate synthase - triose-P sum of glyceraldehyde-3-phosphate and dehydroxyacetone phosphate - UDPGlc uridine diphosphoglucose  相似文献   

16.
D E Hill  G G Hammes 《Biochemistry》1975,14(2):203-213
Equilibrium binding studies of the interaction of rabbit muscle phosphofructokinase with fructose 6-phosphate and fructose 1,6-bisphosphate have been carried out at 5 degrees in the presence of 1-10 mM potassium phosphate (pH 7.0 and 8.0), 5 mM citrate (pH 7.0), or 0.22 mm adenylyl imidodiphosphate (pH 7.0 and 8.0). The binding isotherms for both fructose 6-phosphate and fructose 1,6-bisphosphate exhibit negative cooperativity at pH 7.0 and 8.0 in the presence of 1-10 mM potassium phosphate at protein concentrations where the enzyme exists as a mixture of dimers and tetramers (pH 7.0) or as tetramers (pH 8.0) and at pH 7.0 in the presence of 5 mM citrate where the enzyme exists primarily as dimers. The enzyme binds 1 mol of either fructose phosphate/mol of enzyme monomer (molecular weight 80,000). When enzyme aggregation states smaller than the tetramer are present, the saturation of the enzyme with either ligand is paralleled by polymerization of the enzyme to tetramer, by an increase in enzymatic activity and by a quenching of the protein fluorescence. At protein concentrations where aggregates higher than the tetramer predominate, the fructose 1,6-bisphosphate binding isotherms are hyperbolic. These results can be quantitatively analyzed in terms of a model in which the dimer is associated with extreme negative cooperativity in binding the ligands, the tetramer is associated with less negative cooperativity, and aggregates larger than the tetramer are associated with little or no cooperativity in the binding process. Phosphate is a competitive inhibitor of the fructose phosphate sites at both pH 7.0 and 8.0, while citrate inhibits binding in a complex, noncompetitive manner. In the presence of the ATP analog adenylyl imidodiphosphate, the enzyme-fructose 6-phosphate binding isotherm is sigmoidal at pH 7.0, but hyperbolic at pH 8.0. The characteristic sigmoidal initial velocity-fructose 6-phosphate isotherms for phosphofructokinase at pH 7.0, therefore, are due to an heterotropic interaction between ATP and fructose 6-phosphate binding sites which alters the homotropic interactions between fructose 6-phosphate binding sites. Thus the homotropic interactions between fructose 6-phosphate binding sites can give rise to positive, negative, or no cooperativity depending upon the pH, the aggregation state of the protein, and the metabolic effectors present. The available data suggest the regulation of phosphofructokinase involves a complex interplay between protein polymerization and homotropic and heterotropic interactions between ligand binding sites.  相似文献   

17.
Aspartate transcarbamylase (EC 2.1.3.2) was purified to homogeniety from germinated mung bean seedlings by treatment with carbamyl phosphate. The purified enzyme was a hexamer with a subunit molecular weight of 20,600. The enzyme exhibited multiple activity bands on Polyacrylamide gel electrophoresis, which could be altered by treatment with carbamyl phosphate or UMP indicating that the enzyme was probably undergoing reversible association or dissociation in the presence of these effectors. The carbamyl phosphate stabilized enzyme did not exhibit positive homotropic interactions with carbamyl phosphate and hysteresis. The enzyme which had not been exposed to carbamyl phosphate showed a decrease in specific activity with a change in the concentration of both carbamyl phosphate and protein. The carbamyl phosphate saturation and UMP inhibition patterns were complex with a maximum and a plateau region. The partially purified enzyme also exhibited hysteresis and the hysteretic response, a function of protein concentration, was abolished by preincubation with carbamyl phosphate and enhanced by preincubation with UMP. All these observations are compatible with a postulation that the enzyme activity may be regulated by slow reversible association-dissociation dependent on the interaction with allosteric ligands  相似文献   

18.
The glyceraldehyde 3-phosphate dehydrogenase activity of extracts from heterotrophic Scenedesmus obliquus was linked predominantly to NADH. However, on DEAE-cellulose chromatography the enzyme was eluted by a gradient of phosphate in a form characterized by high NADPH-dependent glyceraldehyde 3-phosphate dehydrogenase activity. This interconversion of enzyme forms could be prevented by the presence of NAD during DEAE-cellulose chromatography.High concentrations of phosphate stimulated the NADPH-dependent activity of the purified enzyme at the expense of activity linked to NADH and these changes were associated with depolymerization of a hexadecamer to a tetramer. The effect of phosphate on the rates of increase in NADPH-dependent activity and of a decrease in activity linked to NADH was cooperative with a Hill coefficient of 3.2. The inversely related changes in coenzyme specificity were inhibited to the same extent by NAD and the response to this ligand was anticooperative. These findings imply a strictly inverse proportional relationship between the rates of change of NADH and NADPH-linked activity. In the presence of dithiothreitol, low concentrations of phosphate promoted NADPH-dependent activity by stabilising the unstable tetrameric form produced from the hexadecamer by the thiol.These phenomena are discussed in relation to a general mechanism for the in vivo promotion of NADPH-dependent glyceraldehyde 3-phosphate dehydrogenase activity.  相似文献   

19.
The effect of pH and of Mg2+ concentration on the light activated form of stromal fructose-1,6-bisphosphatase (FBPase) was studied using the enzyme rapidly extracted from illuminated spinach chloroplasts. The (fructose-1,6-bisphosphate4-)(Mg2+) complex has been identified as the substrate of the enzyme. Therefore, changes of pH and Mg2+ concentrations have an immediate effect on the activity of FBPase by shifting the pH and Mg2+ dependent equilibrium concentration of the substrate. In addition, changes of pH and Mg2+ concentration in the assay medium have a delayed effect on FBPase activity. A correlation of the activities observed using different pH and Mg2+ concentrations indicates, that the effect is not a consequence of the pH and Mg2+ concentration as such, but is caused by a shift in the equilibrium concentration of a hypothetical inhibitor fructose-1,6-bisphosphate3- (uncomplexed), resulting in a change of the activation state of the enzyme. The interplay between a rapid effect on the concentration of the substrate and a delayed effect on the activation state enables a rigid control of stromal FBPase by stromal Mg2+ concentrations and pH. Fructose-1,6-bisphosphatase is allosterically inhibited by fructose-6-phosphate in a sigmoidal fashion, allowing a fine control of the enzyme by its product.Abbreviations Fru1,6 bis P fructose-1,6-bisphosphate - Fru6P fructose-6-phosphate - FBPase fructose-1,6-bisphosphatase Some of these results have been included in a preliminary report (Heldt et al. 1984)  相似文献   

20.
The ability for various ligands to modulate the binding of fructose 1,6-bisphosphate (Fru-1,6-P2) with purified rat liver pyruvate kinase was examined. Binding of Fru-1,6-P2 with pyruvate kinase exhibits positive cooperativity, with maximum binding of 4 mol Fru-1,6-P2 per enzyme tetramer. The Hill coefficient (nH), and the concentration of Fru-1,6-P2 giving half-maximal binding [FBP]1/2, are influenced by several factors. In 150 mM Tris-HCl, 70 mM KCl, 11 mM MgSO4 at pH 7.4, [FBP]1/2 is 2.6 microM and nH is 2.7. Phosphoenolpyruvate and pyruvate enhance the binding of Fru-1,6-P2 by decreasing [FBP]1/2. ADP and ATP alone had little influence on Fru-1,6-P2 binding. However, the nucleotides antagonize the response elicited by pyruvate or phosphoenolpyruvate, suggesting that the competent enzyme substrate complex does not favor Fru-1,6-P2 binding. Phosphorylation of pyruvate kinase or the inclusion of alanine in the medium, two actions which inhibit the enzyme activity, result in diminished binding of low concentrations of Fru-1,6-P2 with the enzyme. These effectors do not alter the maximum binding capacity of the enzyme but rather they raise the concentrations of Fru-1,6-P2 needed for maximum binding. Phosphorylation also decreased the nH for Fru-1,6-P2 binding from 2.7 to 1.7. Pyruvate kinase activity is dependent on a divalent metal ion. Substituting Mn2+ for Mg2+ results in a 60% decrease in the maximum catalytic activity for the enzyme and decreases the concentration of phosphoenolpyruvate needed for half-maximal activity from 1 to 0.1 mM. As a consequence, Mn2+ stimulates activity at subsaturating concentrations of phosphoenolpyruvate, but inhibits at saturating concentrations of the substrate or in the presence of Fru-1,6-P2. Both Mg2+ and Mn2+ diminish binding of low concentrations of Fru-1,6-P2; however, the concentrations of the metal ions needed to influence Fru-1,6-P2 binding exceed those needed to support catalytic activity.  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