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.
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   

2.
袁剑  秦浩  葛向阳  张伟国 《微生物学通报》2011,38(10):1482-1487
L-乳酸脱氢酶(L-lactate dehydrogenase,L-LDH)是发酵生产L-乳酸中催化丙酮酸转化成L-乳酸的关键酶。以干酪乳杆菌G-02(Lactobacillus casei G-02)基因组DNA为模板,克隆得到L-LDH基因(ldhL),经序列分析后将其连接到表达载体pET-28a(+)上,构建成重组质粒pET-ldhL转化到大肠杆菌BL21(DE3)中,实现ldhL基因的表达。30°C加入IPTG诱导表达后,经镍柱亲和层析纯化的重组蛋白样品通过SDS-PAGE分析,约在40 kD处出现显著的特异性条带。对表达的L-LDH生物学特异性研究显示:重组L-LDH的比酶活为1 722 U/mg,最适反应温度为40°C-45°C;果糖-1,6-二磷酸(FBP)为别构激活剂,使最适pH向中性方向偏移(pH为6.6-6.8),Mn2+可拓宽最适酶活pH范围;Mn2+、Ca2+和Mg2+对L-LDH有激活作用,而Zn2+对L-LDH有抑制作用。  相似文献   

3.
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   

4.
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   

5.
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   

6.
The activity of pyruvate kinase of Leishmania mexicana is allosterically regulated by fructose 2,6-bisphosphate (F-2,6-P(2)), contrary to the pyruvate kinases from other eukaryotes that are usually stimulated by fructose 1,6-bisphosphate (F-1,6-P(2)). Based on the comparison of the three-dimensional structure of Saccharomyces cerevisiae pyruvate kinase crystallized with F-1,6-P(2) present at the effector site (R-state) and the L. mexicana enzyme crystallized in the T-state, two residues (Lys453 and His480) were proposed to bind the 2-phospho group of the effector. This hypothesis was tested by site-directed mutagenesis. The allosteric activation by F-2,6-P(2) appeared to be entirely abrogated in the mutated enzymes confirming our predictions.  相似文献   

7.
H P Lerch  R Frank  J Collins 《Gene》1989,83(2):263-270
The gene (L-HicDH) encoding L-2-hydroxyisocaproate dehydrogenase (L-HicDH) from Lactobacillus confusus was cloned in Escherichia coli. A 69-mer oligodeoxyribonucleotide probe, derived to be complementary to the N-terminal amino acid (aa) coding sequence, was used for screening. The complete nucleotide (nt) sequence of the L-HicDH gene was determined. The 5'-end of the mRNA was mapped by primer extension and the promoter identified. Downstream from the L-HicDH gene is a typical Rho-independent terminator. The aa sequence of L-HicDH, deduced from the nt sequence, has an overall similarity of 30% to the aa sequence of L-lactate dehydrogenase (L-LDH) from Lactobacillus casei. The aa residues involved in binding of coenzyme and substrate are highly conserved in L-HicDH with respect to prokaryotic and eukaryotic L-LDHs. The L-HicDH gene could be expressed under control of phage lambda 'Leftward' and 'rightward' promoters in E. coli up to 35% of total cell protein. The enzyme produced under these conditions exhibits full specific activity and is found exclusively in soluble form.  相似文献   

8.
The fructose 1,6-bisphosphate [Fru(1,6)P2]-dependent lactate dehydrogenase in cells of Streptococcus uberis N.C.D.O. 2039 was purified by a procedure that included chromatography on DEAE-cellulose and Blue Sepharose CL-6B in phosphate buffers. The enzyme appeared to interact with Blue Sepharose through NADH-binding sites. The homogeneous enzyme had catalytic properties that were generally similar to those of other Fru(1,6)P2-dependent lactate dehydrogenases, and it had no catalytic activity in the absence of Fru(1,6)P2. Its existence in different forms, depending on conditions, was investigated by ultracentrifugation, analytical gel filtration and activity measurements. It consisted of subunits with Mr 35,900 +/- 500 and, in the presence of adequate concentrations of Fru(1,6)P2, phosphate or NADH, it existed as a tetramer, whereas when these ligands were in lower concentrations or absent, the subunits were in a concentration-dependent association-dissociation equilibrium. Dissociation occurred slowly and inactivated the enzyme, and although added ligands reversed the dissociation, the lost activity was at best only partly restored. An exception occurred when dissociation was caused by a decrease in temperature, in which case the lost activity was fully restored at the original temperature. The tetramer also lost activity at certain ligand concentrations without dissociating. The results together indicated the presence on the enzyme of two classes of binding site for both Fru(1,6)P2 and NADH, and the likelihood that phosphate bound at the same sites as Fru(1,6)P2. Two different ligands together were much more effective at preventing inactivation and dissociation than was expected from their effectiveness when present separately. It was concluded that tetrameric forms of the enzyme rather than the enzyme in association-dissociation equilibrium were involved in the regulation of its activity in vivo.  相似文献   

9.
The involvement of 6-phosphofructo-2-kinase, fructose 2,6-bisphosphate [Fru(2,6)P2] and pyruvate kinase in the insulin-dependent short-term activation of glycolysis was studied in primary cultures of rat hepatocytes. The short-term influence of insulin on these parameters was dependent on the insulin concentration used for the long-term culture. Cells were cultured either with 10 nM or 0.1 nM insulin for 48 h, and are referred to as 'insulin cells' and 'control cells', respectively. Insulin cells exhibited a high level of Fru(2,6)P2. Addition of insulin to insulin cells led to an immediate stimulation of glycolysis (two-fold) and activation of pyruvate kinase. The concentration of Fru(2,6)P2 and activity of 6-phosphofructo-2-kinase remained constant. Control cells exhibited a very low level of Fru(2,6)P2 and low activity of 6-phosphofructo-2-kinase directly after the medium change. However, both parameters increased during a 1-2-h incubation in the absence of insulin. Although the level of Fru(2,6)P2 thus changed up to tenfold the glycolytic rate remained at a constant value. Addition of insulin to control cells led to a 5-8-fold stimulation of glycolysis but only after a 30-90-min lag phase. During this lag period insulin strongly increased sequentially the 6-phosphofructo-2-kinase, the level of Fru(2,6)P2 and the pyruvate kinase activity. The activation of the latter enzyme slightly preceded the onset of the insulin-stimulated glycolysis. Addition of insulin to control cells, which were preincubated for 3 h in the absence of insulin and in which the Fru(2,6)P2 level had risen insulin-independently, led to an immediate increase in glycolysis without a lag phase. It is concluded that in this insulin-sensitive cell system: the changes of glycolytic flux did not correlate with changes in the level of total Fru(2,6)P2 either in insulin or in control cells; an increase in the Fru(2,6)P2 concentration was not obligatory for the insulin-dependent stimulation of glycolysis in insulin cells; activation of pyruvate kinase and thus glycolysis by insulin did not proceed unless the Fru(2,6)P2 level had been elevated above a threshold level. The lack of correlation between total Fru(2,6)P2 levels and the glycolytic flux and the apparent existence of a threshold concentration for Fru(2,6)P2 suggest a permissive action for this effector in enzyme interconversion.  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.
l-Lactate dehydrogenase (l-LDH) of Lactobacillus casei (LCLDH) is a typical bacterial allosteric l-LDH that requires fructose 1,6-bisphosphate (FBP) for its enzyme activity. A mutant LCLDH was designed to introduce an inter-subunit salt bridge network at the Q-axis subunit interface, mimicking Lactobacillus pentosus non-allosteric l-LDH (LPLDH). The mutant LCLDH exhibited high catalytic activity with hyperbolic pyruvate saturation curves independently of FBP, and virtually the equivalent K(m) and V(m) values at pH 5.0 to those of the fully activated wild-type enzyme with FBP, although the K(m) value was slightly improved with FBP or Mn(2+) at pH 7.0. The mutant enzyme exhibited a markedly higher apparent denaturating temperature (T(1/2)) than the wild-type enzyme in the presence of FBP, but showed an even lower T(1/2) without FBP, where it exhibited higher activation enthalpy of inactivation (ΔH(?)). This result is consistent with the fact that the active state is more unstable than the inactive state in allosteric equilibrium of LCLDH. The LPLDH-like network appears to be conserved in many bacterial non-allosteric l-LDHs and dimeric l-malate dehydrogenases, and thus to be a key for the functional divergence of bacterial l-LDHs during evolution.  相似文献   

12.
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   

13.
Bond CJ  Jurica MS  Mesecar A  Stoddard BL 《Biochemistry》2000,39(50):15333-15343
We have analyzed the structural determinants of the allosteric activation of yeast pyruvate kinase (YPK) by mutational and kinetic analysis and initiated a structure-based design project to identify novel effectors that modulate its allosteric response by binding to the allosteric site for fructose-1,6-bisphosphate (FBP). The wild-type enzyme is strongly activated by fructose-1,6-bisphosphate and weakly activated by both fructose-1-phosphate and fructose-6-phosphate; the strength of the activation response is proportional to the affinity of the allosteric effector. A point mutation within the 6'-phosphate binding loop of the allosteric site (T403E) abolishes activation of the enzyme by fructose-1, 6-bisphosphate. The mutant enzyme is also not activated by F1P or F6P. The mutation alone (which incorporates a glutamic acid that is strictly conserved in mammalian M1 isozymes) slightly reduces cooperativity of substrate binding. Three novel compounds were identified that effect the allosteric regulation of YPK by FBP and/or act as novel allosteric activators of the enzyme. One is a physiologically important diphospho sugar, while the other two are hydrophobic compounds that are dissimilar to the natural effector. These results demonstrate that novel allosteric effectors may be identified using structure-based screening and are indicative of the potential of this strategy for drug discovery. Regulatory sites are generally more divergent than catalytic sites and therefore offer excellent opportunities for discrimination and specificity between different organisms or between different tissue types.  相似文献   

14.
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   

15.
The crystal structure of Escherichia coli phosphofructokinase complexed with its reaction products fructose 1,6-bisphosphate (Fru1,6P) and ADP/Mg2+, and the allosteric activator ADP/Mg2+, has been determined at 2.4 A resolution. The structure was solved by molecular replacement using the known structure of Bacillus stearothermophilus phosphofructokinase, and has been refined to a crystallographic R-factor of 0.165 for all data. The crystallization mixture contained the substrate fructose 6-phosphate, but the electron density maps showed clearly the presence of the product fructose 1,6-bisphosphate, presumably formed by the enzyme reaction with contaminating ATP. The crystal consists of tetrameric molecules with subunits in two different conformations despite their chemical identity. The magnesium ion in the "closed" subunit bridges the phosphate groups of the two products. In the "open" subunit, the products are about 1.5 A further apart, with the Mg2+ bound only to ADP. These two conformations probably represent two successive stages along the reaction pathway, in which the closure of the subunit is required to bring the substrates sufficiently close to react. This conformational change within the subunit is distinct from the quaternary structure change seen previously in the inactive T-state conformation. It is probably not involved in the co-operativity or allosteric control of the enzyme, since the co-operative product fructose 1,6-bisphosphate is not moved, nor are the subunit interfaces changed. The structure of the enzyme is similar to that of B. stearothermophilus phosphofructokinase, and confirms the location of the sites for the two reaction products (or substrates), and of the effector site binding the activator ADP/Mg2+. However, this structure gives a clearer picture of the active site, and of the interactions between the enzyme and its reaction products.  相似文献   

16.
The polypeptide chain of the allosteric L-lactate dehydrogenase (EC 1.1.1.27) of Lactobacillus casei consists of 325 amino acid residues. Despite the strikingly different enzymatic characteristics of the allosteric L-lactate dehydrogenase of L. casei and of the non-allosteric vertebrate enzymes, the sequence of the allosteric enzyme shows a distinct homology with that of the non-allosteric vertebrate enzymes (average identity: 37%). An especially high sequence homology can be identified within the active center (average identity: 70%). A clear deviation of the L. casei enzyme from the vertebrate enzyme is the lack of the first 12 amino acid residues at the N terminus and an additional 7 amino acid residues at the C terminus. The localization of the binding site of the allosteric effector D-fructose 1,6-bisphosphate and pH and effector-induced changes of the spectroscopic properties are discussed on the basis of the primary structure.  相似文献   

17.
The allosteric regulation of ADP–glucose pyrophosphorylase is critical for the biosynthesis of glycogen in bacteria and starch in plants. The enzyme from Agrobacterium tumefaciens is activated by fructose 6‐phosphate (Fru6P) and pyruvate (Pyr). The Pyr site has been recently found, but the site where Fru6P binds has remained unknown. We hypothesize that a sulfate ion previously found in the crystal structure reveals a part of the regulatory site mimicking the presence of the phosphoryl moiety of the activator Fru6P. Ser72 interacts with this sulfate ion and, if the hypothesis is correct, Ser72 would affect the interaction with Fru6P and activation of the enzyme. Here, we report structural, binding, and kinetic analysis of Ser72 mutants of the A. tumefaciens ADP‐glucose pyrophosphorylase. By X‐ray crystallography, we found that when Ser72 was replaced by Asp or Glu side chain carboxylates protruded into the sulfate‐binding pocket. They would present a strong steric and electrostatic hindrance to the phosphoryl moiety of Fru6P, while being remote from the Pyr site. In agreement, we found that Fru6P could not activate or bind to S72E or S72D mutants, whereas Pyr was still an effective activator. These mutants also blocked the binding of the inhibitor AMP. This could potentially have biotechnological importance in obtaining enzyme forms insensitive to inhibition. Other mutations in this position (Ala, Cys, and Trp) confirmed the importance of Ser72 in regulation. We propose that the ADP‐glucose pyrophosphorylase from A. tumefaciens have two distinct sites for Fru6P and Pyr working in tandem to regulate glycogen biosynthesis.  相似文献   

18.
Three strains of Lactococcus lactis ssp. lactis, a dairy strain 65.1, a type strain ATCC 19435, and a mutant AS 211, were grown on glucose and on maltose under chemostat conditions. When the culture was shifted from glucose-limiting to non-limiting conditions, the product shifted from mixed acids to lactate. Mixed acids were obtained in all maltose cultures; however, an enhanced lactate formation was observed in 19435 and AS 211. An inorganic-phosphate (Pi)-dependent maltose phosphorylase activity was found to be responsible for the initial conversion of maltose. The activation of maltose phosphorylase by Pi was strain-specific. When growth was on maltose under non-limiting conditions, a correlation was found between high initial maltose phosphorylase and -phosphoglucomutase activities and lactate production. No such correlation was observed in maltose-limited cells. In glucose-grown cells under non-limiting conditions, homo-fermentative lactate formation coincided with high concentrations of fructose 1,6-bisphosphate (Fru1,6P 2) and pyruvate (Pyr) and low concentrations of phosphoenolpyruvate (PPyr). Under limiting conditions, mixed acid formation coincided with low concentrations of Fru1,6P 2 and Pyr and high concentrations of PPyr. In maltose-grown cells there was no correlation between intracellular intermediary metabolite concentrations and product formation. Therefore, in addition to intracellular intermediary metabolite concentrations, the product formation on maltose is suggested to be regulated by the transport and initial phosphorylating steps.  相似文献   

19.
Purpose

The aim of this study was to prove that Terrilactibacillus laevilacticus SK5-6, a novel D-lactate producer, exhibited a good fermentation performance comparing to the reference D-lactate producer Sporolactobacillus sp.

Methods

Glucose bioconversion for D-lactate production and the activity of five key enzymes including phosphofructokinase (PFK), pyruvate kinase (PYK), D-lactate dehydrogenase (D-LDH), L-lactate dehydrogenase (L-LDH), and lactate isomerase (LI) were investigated in the cultivation of T. laevilacticus SK5–6 and S. laevolacticus 0361T.

Results

T. laevilacticus SK5–6 produced D-lactate at higher yield, productivity, and optical purity compared with S. laevolacticus 0361T. T. laevilacticus SK5–6, the catalase-positive isolate, simultaneously grew and produced D-lactate without lag phase while delayed growth and D-lactate production were observed in the culture of S. laevolacticus 0361T. The higher production of D-lactate in T. laevilacticus SK5–6 was due to the higher growth rate and the higher specific activities of the key enzymes observed at the early stage of the fermentation. The low isomerization activity was responsible for the high optical purity of D-lactate in the cultivation of T. laevilacticus SK5–6.

Conclusion

The lowest specific activity of PFK following by PYK and D/L-LDHs, respectively, indicated that the conversion of fructose-6-phosphate was the rate limiting step. Under the well-optimized conditions, the activation of D/L-LDHs by fructose-1,6-phosphate and ATP regeneration by PYK drove glucose bioconversion toward D-lactate. The optical purity of D-lactate was controlled by D/L-LDHs and the activation of isomerases. High D-LDH with limited isomerase activity was preferable during the fermentation as it assured the high optical purity.

  相似文献   

20.
Metabolic control of hepatic gluconeogenesis during exercise.   总被引:2,自引:0,他引:2       下载免费PDF全文
Prolonged exercise increased the concentrations of the hexose phosphates and phosphoenolpyruvate and depressed those of fructose 1,6-bisphosphate, triose phosphates and pyruvate in the liver of the rat. Since exercise increases gluconeogenic flux, these changes in metabolite concentrations suggest that metabolic control is exerted, at least, at the fructose 6-phosphate/fructose 1,6-bisphosphate and phosphoenolpyruvate/pyruvate substrate cycles. Exercise increased the maximal activities of glucose 6-phosphatase, fructose 1,6-bisphosphatase, pyruvate kinase and pyruvate carboxylase in the liver, but there were no changes in those of glucokinase, 6-phosphofructokinase and phosphoenolpyruvate carboxykinase. Exercise changed the concentrations of several allosteric effectors of the glycolytic or gluconeogenic enzymes in liver; the concentrations of acetyl-CoA, ADP and AMP were increased, whereas those of ATP, fructose 1,6-bisphosphate and fructose 2,6-bisphosphate were decreased. The effect of exercise on the phosphorylation-dephosphorylation state of pyruvate kinase was investigated by measuring the activities under conditions of saturating and subsaturating concentrations of substrate. The submaximal activity of pyruvate kinase (0.5 mM-phosphoenolpyruvate), expressed as percentage of Vmax., decreased in the exercised animals to less than half that found in the controls. These changes suggest that hepatic pyruvate kinase is less active during exercise, possibly owing to phosphorylation of the enzyme, and this may play a role in increasing the rate of gluconeogenesis.  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