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 kinetic mechanisms of Escherichia coli phosphofructokinase-2 (Pfk-2) and of the mutant enzyme Pfk-2 were investigated. Initial velocity studies showed that both enzymes have a sequential kinetic mechanism, indicating that both substrates must bind to the enzyme before any products are released. For Pfk-2, the product inhibition kinetics was as follows: fructose-1,6-P2 was a competitive inhibitor versus fructose-6-P at two ATP concentrations (0.1 and 0.4 mM), and noncompetitive versus ATP. The other product inhibition patterns, ADP versus either ATP or fructose-6-P were noncompetitive. Dead-end inhibition studies with an ATP analogue, adenylyl imidodiphosphate, showed uncompetitive inhibition when fructose-6-P was the varied substrate. For Pfk-2, the product inhibition studies revealed that ADP was a competitive inhibitor versus ATP at two fructose-6-P concentrations (0.05 and 0.5 mM), and noncompetitive versus fructose-6-P. The other product, fructose-1, 6-P2, showed noncompetitive inhibition versus both substrates, ATP and fructose-6-P. Sorbitol-6-P, a dead-end inhibitor, exhibited competitive inhibition versus fructose-6-P and uncompetitive versus ATP. These results are in accordance with an Ordered Bi Bi reaction mechanism for both enzymes. In the case of Pfk-2, fructose-6-P would be the first substrate to bind to the enzyme, and fructose-1,6-P2 the last product to be released. For Pfk-2, ATP would be the first substrate to bind to the enzyme, and APD the last product to be released.  相似文献   

2.
Arg252 of fructose-6-phosphate 1-kinase (PFK) from Bacillus stearothermophilus has been proposed to be involved in the binding of the substrate Fru-6-P. We demonstrate here that mutation of this residue to alanine converts the enzyme to a form with characteristics similar to those of its allosterically tight form. The mutant enzyme exhibits a high affinity for its inhibitor phosphoenolpyruvate (a 68-fold difference compared to wild type) and a dramatically decreased Fru-6-P affinity (1500-fold increase in Km). It is more sensitive to inhibition by high ATP concentrations than the wild type, and this inhibition is relieved by ADP, GDP, or higher Fru-6-P concentrations. In contrast, mutation of Arg252 to lysine increases the affinity of the enzyme for P-enolpyruvate by only 2-fold and increases its Km for Fru-6-P by only 50-fold. Sigmoidal kinetics with respect to Fru-6-P in the presence of P-enolpyruvate were observed with Hill numbers of 2.2, 2.4, and 1.7 for wild-type B. stearothermophilus PFK and the Arg252 to lysine and to alanine mutations, respectively. Unlike fructose-6-phosphate 1-kinase from Escherichia coli, in the absence of P-enolpyruvate, B. stearothermophilus PFK exhibits a hyperbolic profile with respect to Fru-6-P concentration. B. stearothermophilus PFK is sensitive to inhibition by high ATP concentrations and competitively inhibited by GDP or ADP. Our data indicate that Arg252 of B. stearothermophilus PFK plays a major role in both Fru-6-P binding and allosteric interaction between the subunits. However, this residue does not seem to participate directly in the catalytic process.  相似文献   

3.
The effect of natural "activation factor" and synthetic fructose-2,6-P2 on the allosteric kinetic properties of liver and muscle phosphofructokinases was investigated. Both synthetic and natural fructose-2,6-P2 show identical effects on the allosteric kinetic properties of both enzymes. Fructose-2,6-P2 counteracts inhibition by ATP and citrate and decreases the Km for fructose-6-P. This fructose ester also acts synergistically with AMP in releasing ATP inhibition. The Km values of liver and muscle phosphofructokinase for fructose-2,6-P2 in the presence of 1.25 mM ATP are 12 milliunits/ml (or 24 nM) and 5 milliunits/ml (or 10 nM), respectively. At near physiological concentrations of ATP (3 mM) and fructose-6-P (0.2 mM), however, the Km values for fructose-2,6-P2 are increased to 12 microM and 0.8 microM for liver and muscle enzymes, respectively. Thus, fructose-2,6-P2 is the most potent activator of the enzyme compared to other known activators such as fructose-1,6-P2. The rates of the reaction catalyzed by the enzymes under the above conditions are nonlinear: the rates decelerate in the absence or in the presence of lower concentrations of fructose-2,6-P2, but the rates become linear in the presence of higher concentrations of fructose-2,6-P2. Fructose-2,6-P2 also protects phosphofructokinase against inactivation by heat. Fructose-2,6-P2, therefore, may be the most important allosteric effector in regulation of phosphofructokinase in liver as well as in other tissues.  相似文献   

4.
Fenton AW  Reinhart GD 《Biochemistry》2003,42(43):12676-12681
Phosphofructokinase from Escherichia coli (EcPFK) is a homotetramer with four active sites, which bind the substrates fructose-6-phosphate (Fru-6-P) and MgATP. In the presence of low concentrations of Fru-6-P, MgATP displays substrate inhibition. Previous proposals to explain this substrate inhibition have included both kinetic and allosteric mechanisms. We have isolated hybrid tetramers containing one wild type subunit and three mutated subunits (1:3). The mutated subunits contain mutations that decrease affinity for Fru-6-P (R243E) or MgATP (F76A/R77D/R82A) allowing us to systematically simplify the possible allosteric interactions between the two substrates. In the absence of a rate equation to explain the allosteric effects in a tetramer, the data have been compared to simulated data for an allosteric dimer. Since the apparent substrate inhibition caused by MgATP binding is not seen in hybrid tetramers with only a single native MgATP binding site, the proposed kinetic mechanism is not able to explain this phenomenon. The data presented are consistent with an allosteric antagonism between MgATP in one active site and Fru-6-P in a second active site.  相似文献   

5.
Wang X  Kemp RG 《Biochemistry》1999,38(14):4313-4318
The apparent affinity of phosphofructo-1-kinase (PFK) of Escherichia coli for ATP is at least 10 times higher than for other nucleotides. Mutagenesis was directed toward five residues that may interact with ATP: Y41, F76, R77, R82, and R111. Alanine at position 41 or 76 increased the apparent Km by 49- and 62-fold, respectively. Position 41 requires the presence of a large hydrophobic residue and is not restricted to aromatic rings. Tryptophan and, to a lesser extent, phenylalanine could substitute at position 76. None of the mutants at 41 or 76 showed a change in the preference for alternative purines, although F76W used CTP 3 times better than the wild type enzyme. Mutations of R77 suggested that the interaction was hydrophobic with no influence on nucleotide preference. Mutation of R82 to alanine or glutamic acid increased the apparent Km for ATP by more than 20-fold and lowered the kcat/Km with ATP more than 30-fold. However, these mutants had a higher kcat/Km than wild type for both GTP and CTP, reflecting a loss of substrate preference. A loss in preference is seen as well with R111A where the kcat/Km for ATP decreases by only 68%, but the kcat/Km with GTP increases more than 10-fold. Activities with ITP, CTP, and UTP are also higher than with the wild type enzyme. Arginine residues at positions 82 and 111 are important dictators of nucleoside triphosphate preference.  相似文献   

6.
Grapefruit juice sac ATP-PFK was studied kinetically for its substrates ATP and Fru-6-P at pH = 7.5. The Km for ATP is equal to 39.8 +/- 4.6 microM. ATP becomes inhibitory at concentrations above 80 microM. The Km for ATP is not affected by the addition of citrate (10 mM). For Fru-6-P, the saturation curve is sigmoidal, with an S0.5 equal to 0.17 +/- 0.03 mM, in the presence of Mg++ (2.5 mM) and ATP (1 mM). ATP-PFK shows a negative cooperativity at lower concentrations of Fru-6-P (h = 0.5), while higher concentrations of the substrate induce a positive cooperation (h = 1.5). The presence of citrate affects the S0.5 affinity value, but not the Vmax. The presence of citrate (10 mM) removes the cooperative effect at higher concentrations of the substrate, as h = 1.0. A theoretical Ki for citrate was calculated and equals 1.30 mM.  相似文献   

7.
Two approaches have been used to study the allosteric modulation of phosphofructokinase at physiological concentration of enzyme; a "slow motion" approach based on the use of a very low Mg2+/ATP ratio to conveniently lower Vmax, and the addition of polyethylene glycol as a "crowding" agent to favor aggregation of diluted enzyme. At 0.6 mg/ml muscle phosphofructokinase exhibited a drastic decrease in the ATP inhibition and the concomitant increase in the apparent affinity for fructose-6-P, as compared to a 100-fold diluted enzyme. Similar results were obtained with diluted enzyme in the presence of 10% polyethylene glycol (Mr = 6000). Results with these two approaches in vitro were essentially similar to those previously observed in situ (Aragón, J. J., Felíu, F. E., Frenkel, R., and Sols, A. (1980) Proc. Natl. Acad. Sci. U. S. A. 77, 6324-6328), indicating that the enzyme is strongly dependent on homologous interactions at physiological concentrations. With polyethylene glycol it was observed that within the physiological range of concentration of substrates and the other positive effectors, fructose-2,6-P2 still activates the liver phosphofructokinase although it no longer significantly affects the muscle isozyme. In the presence of polyethylene glycol, muscle phosphofructokinase can approach its maximal rate even in the presence of physiologically high concentrations of ATP. Three minor activities of muscle phosphofructokinase have been studied at high enzyme concentration: the hydrolysis of MgATP (ATPase) and fructose-1,6-P2 (FBPase), produced in the absence of the other substrate, and the reverse reaction from MgADP and fructose-1,6-P2. The kinetic study of these activities has allowed a new insight into the mechanisms involved in the modulation of phosphofructokinase activity. The binding of (Mg)ATP at its regulatory site reduces the ability of the enzyme to cleave the bond of the terminal phosphate of MgATP at the substrate site. The positive effectors (Pi, cAMP, NH+4, fructose-1,6-P2, and fructose-2,6-P2) decrease the inhibitory effect of MgATP. Citrate and fructose-2,6-P2 both act as mechanistically "secondary" effectors in the sense that citrate does not inhibit and fructose-2,6-P2 does not activate the FBPase activity, requiring both the presence of ATP to affect the enzyme activity. In conclusion it appears that the regulatory behavior of mammalian phosphofructokinases is utterly dependent on the fact of their high concentrations in vivo.  相似文献   

8.
The 1-phosphofructokinase (1-PFK, EC 2.7.1.56) from Pseudomonas putida was partially purified by a combination of (NH4)2SO4 fractionation and DEAE-Sephadex column chromatography. In its kinetic properties, this enzyme resembled the 1-PFK's from other bacteria. With the substrates fructose-1-phosphate (F-1-P) and adenosine triphosphate (ATP) Michaelis-Menten kinetics were observed, the Km for one substrate being unaffected by a variation in the concentration of the other substrate. At pH 8.0, the Km values for F-1-P and ATP were 1.64 X 10(-4) M and 4.08 X 10(-4) M, respectively. At fixed concentrations of F-1-P and ATP, an increase in the Mg2+ resulted in sigmoidal kinetics. Activity was inhibited by ATP when the ratio of ATP:Mg2+ was greater than 0.5 suggesting that ATP:2 Mg2+ was the substrate and free ATP was inhibitory. Activity of 1-PFK was stimulated by K+ and to a lesser extent by NH4+ and Na+. The reaction rate was unaffected by 2 mM K2HPO4, pyruvate, phosphoenolpyruvate, adenosine monophosphate, adenosine 3',5'-cyclic monophosphate, fructose-6-phosphate, glucose-6-phosphate, 6-phosphogluconate, 2-keto-3-deoxy-6-phosphogluconate, or citrate. The results indicated that the 1-PFK from P. putida was not allosterically regulated by a number of metabolites which may play an important role in the catabolism of D-fructose.  相似文献   

9.
The activity of yeast phosphofructokinase assayed in vitro at physiological concentrations of known substrates and effectors is 100-fold lower than the glycolytic flux observed in vivo. Phosphate synergistically with AMP activates the enzyme to a level within the range of the physiological needs. The activation by phosphate is pH-dependent: the activation is 100-fold at pH 6.4 while no effect is observed at pH 7.5. The activation by AMP, phosphate, or both together is primarily due to changes in the affinity of the enzyme for fructose-6-P. Under conditions similar to those prevailing in glycolysing yeast (pH 6.4, 1 mM ATP, 10 mM NH4+) the apparent affinity constant for fructose-6-P (S0.5) decreases from 3 to 1.4 mM upon addition of 1 mM AMP or 10 mM phosphate; if both activators are present together, S0.5 is further decreased to 0.2 mM. In all cases the cooperativity toward fructose-6-P remains unchanged. These results are consistent with a model for phosphofructokinase where two conformations, with different affinities for fructose-6-P and ATP, will present the same affinity for AMP and phosphate. AMP would diminish the affinity for ATP at the regulatory site and phosphate would increase the affinity for fructose-6-P. The results obtained indicate that the activity of phosphofructokinase in the shift glycolysis-gluconeogenesis is mainly regulated by changes in the concentration of fructose-6-P.  相似文献   

10.
The interaction of several inhibitors with muscle phosphofructokinase has been studied by both equilibrium binding measurements and kinetic analysis. At low concentrations of citrate a maximum of 1 mol is bound per mol of enzyme protomer. Tight binding requires MgATP and very weak binding is observed in the absence of either magnesium ion or ATP. ITP at low concentrations cannot replace ATP. In the presence of MgATP and at pH 7.0, the dissociation constant for the enzyme-citrate complex is 20 muM. At 50 muM citrate and excess magnesium ion, the concentration of ATP required to give half-maximal binding of citrate is approximately 3 muM . Both P-enolpyruvate and 3-P-glycerate compete for the binding of citrate and the estimated Ki values are 480 and 52 muM, respectively. Creatine-P, another inhibitor of muscle phosphofructokinase, does not compete with the binding of citrate. Measurement of the equilibrium binding of ATP shows that citrate, 3-P-glycerate, P-enolpyruvate, and creatine-P all increase the affinity of enzyme for MgATP with the concentration required to give an effect increasing in the order given. In kinetic studies, citrate, 3-P-glycerate and P-enolpyruvate each act synergistically with ATP to inhibit the phosphofructokinase reaction. This is indicated by the observation that the three metabolites do not inhibit the enzyme with ITP as the phosphoryl donor and that they inhibit at ATP concentrations that are not themselves inhibitory. Furthermore, the sensitivity to the inhibitors increases with increasing ATP concentrations. Striking differences in the extent of inhibition can be seen by varying the order of addition of assay components. Preincubation of the enzyme with ATP and citrate, 3-P-glycerate, or P-enolpyruvate results in greater inhibition than when the inhibitor is added after the reaction is started with fructose-6-P. Furthermore, the inhibition is reversed partially 10 to 15 min after the addition of fructose-6-P. This phenomenon is particularly striking with creatine-P as the inhibitor. Very high concentrations of this inhibitor are required to show any effect if the inhibitor is added after fructose-6-P. These effects are interpreted as reflecting slow conformational changes between an active form with high affinity for fructose-6-P and an inactive, or less active, conformation that binds the inhibitors. Citrate, 3-P-glycerate, P-enolpyruvate, and creatine-P increase the rate of the phosphofructokinase at subsaturating concentrations of MgITP. The results indicate a common binding site on the enzyme for citrate, 3-P-glycerate, and P-enolpyruvate that is distinct from the ATP inhibitory site. An additional site (or sites) for creatine-P is indicated. All four inhibitors act synergistically with ATP by increasing the affinity of the enzyme for MgATP at an inhibitory site. The inhibitors appear also to increase the affinity of the catalytic nucleoside triphosphate site for substrate.  相似文献   

11.
Partially purified fructose-6-P,2-kinase:fructose-2,6-bisphosphatase from beef heart was phosphorylated by cAMP protein kinase. The phosphorylated fructose-6-P,2-kinase shows lower Km for Fru-6-P (43 versus 105 microM) and for ATP (0.55 versus 1.3 mM) but no change in the Vmax, compared to those for unphosphorylated enzyme. There was no detectable change in Km or Vmax of fructose-2,6-bisphosphatase activity by the phosphorylation. These changes in heart fructose-6-P,2-kinase were in direct contrast to previous results for the liver isozyme in which phosphorylation led to inhibition of the kinase activity and activation of the phosphatase activity.  相似文献   

12.
Stopped-flow measurements have been carried out to study some basic allosteric properties of muscle and yeast phosphofructokinase at physiological concentration of enzyme. An important increase in the affinity for fructose-6-P accompanied by an intense decrease in the ATP inhibition was observed with the muscle enzyme, which also became insensitive to fructose-2,6-P2 under these conditions. Yeast phosphofructokinase exhibited a significant diminution in the inhibition by ATP, although with no apparent change in the affinity for fructose-6-P. These results provide strong support in favor of the dependence of the allosteric regulation of phosphofructokinase on its concentration in vivo.  相似文献   

13.
In the presence of its allosteric activator GDP, the major phosphofructokinase-1 from Escherichia coli K12 follows Michaelis—Menten kinetics. The kinetic behavior observed at steady-state using different concentrations of the substrates ATP and fructose-6-phosphate and the pattern of inhibition by the substrate analogs adenylyl-(β,γ-methylene)-diphosphonate and D-arabinose-5-phosphate are consistent with a random sequential mechanism in rapid equilibrium, rather than with an ordered binding as was suggested earlier. However, ATP and fructose-6-phosphate do not bind independently to the same active site, since the apparent affinity for one substrate is decreased about 20-fold when the other substrate is already bound. The antagonism between ATP and fructose-6-phosphate shows that a negative interaction occurs during the reaction with E. coli phosphofructokinase-1 which must be considered in addition to its allosteric properties.  相似文献   

14.
The kinetic properties of 6-phosphofructo-1-kinase (PFK) from skeletal muscle (PFKM) of gilthead sea bream (Sparus aurata) were studied, after 10,900-fold purification to homogeneity. The native enzyme had an apparent molecular mass of 662 kDa and is composed of 81 kDa subunits, suggesting a homooctameric structure. At physiological pH, S. aurata PFKM exhibited sigmoidal kinetics for the substrates, fructose-6-phosphate (fru-6-P) and ATP. Fructose-2,6-bisphosphate (fru-2,6-P(2)) converted the saturation curves for fru-6-P to hyperbolic, activated PFKM synergistically with other positive effectors of the enzyme such as AMP and ADP, and counteracted ATP and citrate inhibition. The fish enzyme showed differences regarding other animal PFKs: it is active as a homooctamer, and fru-2,6-P(2) and pH affected affinity for ATP. By monitoring incorporation of (32)P from ATP, we show that fish PFKM is a substrate for the cAMP-dependent protein kinase. The mechanism involved in PFKM activation by phosphorylation contrasts with previous observations in other species: it increased V(max) and did not affect affinity for fru-6-P. Unlike the mammalian muscle enzyme, our findings support that phosphorylation of PFKM may exert a major role during starvation in fish muscle.  相似文献   

15.
Inactivation of a bifunctional enzyme, fructose-6-P,2-kinase:fructose-2,6-bisphosphatase by pyridoxal 5'-P followed by reduction with NaBH4 was studied. Fructose-6-P,2-kinase is over 80% inactivated by 2 mM pyridoxal 5'-P. The stoichiometry of the pyridoxyl-P incorporation and the inactivation of the kinase follows a biphasic curve. The first P-pyridoxyl residue incorporated per protomer does not affect fructose-6-P,2-kinase, but the next two P-pyridoxyl incorporation/protomer results in 80% inactivation. The Km values for ATP and fructose-6-P of the enzymes containing varying amounts of P-pyridoxyl groups at intermediate levels of inactivation are not altered, but Vmax is decreased. Among the metabolites tested, only fructose-2,6-P2 and Mg-ATP are competitive with pyridoxal-P and protect the enzyme against the inactivation. Neither the activity nor the fructose-6-P inhibition of fructose-2,6-bisphosphatase is affected by the modification. The acid hydrolysate of the inactive P-[3H]pyridoxyl enzyme contained only [3H]pyridoxyl lysine. High performance liquid chromatography of tryptic peptides of phospho[3H]pyridoxyl enzymes reveals two peptides which were missing in the enzyme protected by fructose-2,6-P2 or ATP during the modification reaction. These peptides have been isolated, and their amino acid sequences have been determined as Asp-Gln-Asp-Lys-Tyr-Arg and Asp-Val-His-Lys-Tyr. Pyridoxal-P reacts specifically with two lysine residues at the fructose-2,6-P2-binding site of fructose-6-P,2-kinase but not that of fructose-2,6-bisphosphatase. The site may also overlap with the ATP-binding site.  相似文献   

16.
Isolated rat hepatocytes convert 2,5-anhydromannitol to 2,5-anhydromannitol-1-P and 2,5-anhydromannitol-1,6-P2. Cellular concentrations of the monophosphate and bisphosphate are proportional to the concentration of 2,5-anhydromannitol and are decreased by gluconeogenic substrates but not by glucose. Rat liver phosphofructokinase-1 phosphorylates 2,5-anhydromannitol-1-P; the rate is less than that for fructose-6-P but is stimulated by fructose-2,6-P2. At 1 mM fructose-6-P, bisphosphate compounds activate rat liver phosphofructokinase-1 in the following order of effectiveness: fructose-2,6-P2 much greater than 2,5-anhydromannitol-1,6-P2 greater than fructose-1,6-P2 greater than 2,5-anhydroglucitol-1,6-P2. High concentrations of fructose-1,6-P2 or 2,5-anhydromannitol-1,6-P2 inhibit phosphofructokinase-1. Rat liver fructose 1,6-bisphosphatase is inhibited competitively by 2,5-anhydromannitol-1,6-P2 and noncompetitively by 2,5-anhydroglucitol-1,6-P2. The AMP inhibition of fructose 1,6-bisphosphatase is potentiated by 2,5-anhydroglucitol-1,6-P2 but not by 2,5-anhydromannitol-1,6-P2. Rat liver pyruvate kinase is stimulated by micromolar concentrations of 2,5-anhydromannitol-1,6-P2; the maximal activation is the same as for fructose-1,6-P2. 2,5-Anhydroglucitol-1,6-P2 is a weak activator. 2,5-Anhydromannitol-1-P stimulates pyruvate kinase more effectively than fructose-1-P. Effects of glucagon on pyruvate kinase are not altered by prior treatment of hepatocytes with 2,5-anhydromannitol. Pyruvate kinase from glucagon-treated hepatocytes has the same activity as the control pyruvate kinase at saturating concentrations of 2,5-anhydromannitol-1,6-P2 but has a decreased affinity for 2,5-anhydromannitol-1,6-P2 and is not stimulated by 2,5-anhydromannitol-1-P. The inhibition of gluconeogenesis and enhancement of glycolysis from gluconeogenic precursors in hepatocytes treated with 2,5-anhydromannitol can be explained by an inhibition of fructose 1,6-bisphosphatase, an activation of pyruvate kinase, and an abolition of the influence of phosphorylation on pyruvate kinase.  相似文献   

17.
It has been found that the inhibition of Dictyostelium discoideum fructose-1,6-bisphosphatase by fructose 2,6-P2 greatly diminished when the pH was raised to the range 8.5-9.5, which resulted in a marked decrease of the affinity for the inhibitor with no change in the Km for the substrate. This provides evidence for the involvement of an allosteric site for fructose 2,6-P2. Moreover, the fact that excess substrate inhibition also decreased at the pH values for minimal fructose 2,6-P2 inhibition, and was essentially abolished in the presence of fructose 2,6-P2, strongly suggests that this inhibition takes place by binding of fructose 1,6-P2 as a weak analogue of the physiological effector fructose 2,6-P2.  相似文献   

18.
In the presence of its allosteric activator GDP, the major phosphofructokinase-1 from Escherichia coli K12 follows Michaelis—Menten kinetics. The kinetic behavior observed at steady-state using different concentrations of the substrates ATP and fructose-6-phosphate and the pattern of inhibition by the substrate analogs adenylyl-(β,γ-methylene)-diphosphonate and D-arabinose-5-phosphate are consistent with a random sequential mechanism in rapid equilibrium, rather than with an ordered binding as was suggested earlier. However, ATP and fructose-6-phosphate do not bind independently to the same active site, since the apparent affinity for one substrate is decreased about 20-fold when the other substrate is already bound. The antagonism between ATP and fructose-6-phosphate shows that a negative interaction occurs during the reaction with E. coli phosphofructokinase-1 which must be considered in addition to its allosteric properties.  相似文献   

19.
This study is an attempt to relate in vivo use of the 6-phosphogluconate dehydrogenase reaction in Escherichia coli with the characteristics of the enzyme determined in vitro. 1) The enzyme was obtained pure by affinity chromatography and kinetically characterized; as already known, ATP and fructose-1,6-P2 were inhibitors. 2) A series of isogenic strains were made in which in vivo use of thereaction might differ, e.g. a wild type strain versus a mutant lacking 6-phosphogluconate dehydrase, as grown on gluconate; a phosphoglucose isomerase mutant grown on glucose or glycerol. 3) The in vivo rate of use of the 6-phosphogluconate dehydrogenase reaction was determined from measurements of growth rate and yield and from the specific activity of alanine after growth in 1-14C-labeled substrates. 4) The intracellular concentrations of 6-phosphogluconate, NADP+, fructose-1,6-P2, and ATP were measured for the strains in growth on several carbon sources. 5) The metabolite concentrations were used for assay of the enzyme in vitro. The results allow one to calculate how fast the reaction would function in vivo if ATP and fructose-1,6-P2 were its important effectors and if the in vitro assay conditions apply in vivo. The predicted in vivo rates ranged down to as low as one-tenth of the actual rates, and, accordingly, one cannot yet draw firm conclusions about how the reaction is actually controlled in vivo.  相似文献   

20.
Twenty-five metabolites of glucose, gluconeogenic substrates, and related compounds were examined as potential inhibitors of glucose-6-phosphatase (EC 3.1.3.9) catalytic unit and substrate transport function, using disrupted and intact rat liver microsomes. Inhibitions (competitive) were noted with six. Calculated per cent inhibitions with presumed near-physiologic concentrations of inhibitor and substrate were small. However, when hepatic fructose-1-P concentration is elevated in response to a fructose load, inhibition of glucose-6-phosphatase by fructose-1-P may play a regulatory role, along with fructose-1-P-associated deinhibition of glucokinase, by directing glucose-6-P away from glucose formation and towards glycogen synthesis and glycolysis.  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