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 B-type cofactor-dependent phosphoglycerate mutase (dPGM-B) catalyzes the interconversion of 2-phosphoglycerate and 3-phosphoglycerate in glycolysis and gluconeogenesis pathways using 2,3-bisphosphoglycerate as the cofactor. The crystal structures of human dPGM-B bound with citrate were determined in two crystal forms. These structures reveal a dimerization mode conserved in both of dPGM and BPGM (bisphosphoglycerate mutase), based on which a dPGM/BPGM heterodimer structure is proposed. Structural comparison supports that the conformational changes of residues 13-21 and 98-117 determine PGM/BPGM activity differences. The citrate-binding mode suggests a substrate-binding model, consistent with the structure of Escherichia coli dPGM/vanadate complex. A chloride ion was found in the center of the dimer, providing explanation for the contribution of chloride ion to dPGM activities. Based on the structural information, the possible reasons for the deficient human dPGM mutations found in some patients are also discussed.  相似文献   

2.
Rat bone marrow cells have been fractionated by density gradient in Percoll. Differential counting of erythroid cells, haemoglobin concentration and bisphosphoglycerate mutase and phosphoglycerate kinase activities have been determined in cellular fractions. As shown by means of a statistical approach, an increase in bisphosphoglycerate mutase activity and a slight decrease in phosphoglycerate kinase activity is found in erythroid cells as their haemoglobin content increases. Our results suggest that there is a synthesis of 2,3-bisphosphoglycerate during the erythropoietic process which parallels the synthesis of haemoglobin.  相似文献   

3.
Cloning and sequencing of a murine cDNA with the entire coding region of 2,3-bisphosphoglycerate mutase is reported, as a prerequisite for further expression studies of this erythroid specific enzyme in Friend mouse erythroleukemia cells. A comparison between species of the deduced amino acid sequences of these proteins shows 20 substitutions between mouse and human and 21 between mouse and rabbit: none of these substitutions are in positions assumed to be in the active site. Amino acid alignment with the other related enzymes, the phosphoglycerate mutases, in combination with crystallographic data from yeast phosphoglycerate mutase, gives some insight into the structure/function correlation for this protein family. Amino acid residues which are most likely critical for either 2,3-bisphosphoglycerate mutase or phosphoglycerate mutase function are pointed out. Concerning the phylogenetic analysis, phosphoglycerate mutases B and M from mammalians appear to have diverged with the yeast enzyme from a common ancestor, before the emergence of the 2,3-bisphosphoglycerate mutases.  相似文献   

4.
The bisphosphatase domain of the rat liver 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase has been shown to exhibit a structural similarity to yeast phosphoglycerate mutase and human red blood cell 2,3-bisphosphoglycerate mutase including very similar active site sequences with a histidyl residue being involved in phospho group transfer. The liver bifunctional enzyme was found to catalyze the hydrolysis of glycerate 1,3-bisphosphate to glycerate 3-phosphate and inorganic phosphate. The Km for glycerate 1,3-bisphosphate was 320 microM and the Vmax was 11.5 milliunits/mg. Incubation of the rat liver enzyme with [1-32P]glycerate 1,3-bisphosphate resulted in the formation of a phosphoenzyme intermediate, and the labeled amino acid was identified as 3-phosphohistidine. Tryptic and endoproteinase Lys-C peptide maps of the 32P-phosphoenzyme labeled either with [2-32P]fructose 2,6-bisphosphate or [1-32P]glycerate 1,3-bisphosphate revealed that 32P-radioactivity was found in the same peptide, proving that the same histidyl group accepts phosphate from both substrates. Fructose 2,6-bisphosphate inhibited competitively the formation of phosphoenzyme from [1-32P]glycerate 1,3-bisphosphate. Effectors of fructose-2,6-bisphosphatase also inhibited phosphoenzyme formation. Substrates and products of phosphoglycerate mutase and 2,3-bisphosphoglycerate mutase also modulated the activities of the bifunctional enzyme. These results demonstrate that, in addition to a structural homology, the bisphosphatase domain of the bifunctional enzyme has a functional similarity to phosphoglycerate mutase and 2,3-bisphosphoglycerate mutase and support the concept of an evolutionary relationship between the three enzyme activities.  相似文献   

5.
Two enzymes which possess 2,3-bisphosphoglycerate synthase, 2,3-bisphosphoglycerate phosphatase and phosphoglycerate mutase activities have been purified from pig skeletal muscle. One of the enzymes corresponds to type M phosphoglycerate mutase. The other enzyme shows properties similar to those of the 2,3-bisphosphoglycerate synthase-phosphatase present in mammalian erythrocytes. The erythrocyte and the muscle enzyme possess the same molecular (56 000) and subunit (27 000) weights. The synthase, phosphatase and mutase activity ratio is similar in both enzymes, and they are affected by the same inhibitor (glycerate 3-P) and activators (glycolate 2-P, pyrophosphate, sulfite and bisulfite).  相似文献   

6.
Type M phosphoglycerate mutase and skeletal muscle bisphosphoglycerate synthase-phosphatase from pig are similarly affected by Hg2+. Both enzymes lose the phosphoglycerate mutase and the glycerate-2,3-P2 synthase activities, and increase the glycerate-2,3-P2 phosphatase activity upon Hg2+-treatment. In contrast, bisphosphoglycerate phosphatase from pig skeletal muscle is inactivated by Hg2+. These results confirm the similarity between phosphoglycerate mutase and bisphosphoglycerate synthase-phosphatase. In addition they support the existence of separate binding sites for monophosphoglycerates and for bisphosphoglycerates at the phosphoglycerate mutase active site.  相似文献   

7.
2,3-Bisphosphoglycerate synthase-phosphatase and the hybrid phosphoglycerate mutase/2,3-bisphosphoglycerate synthase-phosphatase have been partially purified from pig brain. Their 2,3-bisphosphoglycerate synthase, 2,3-bisphosphoglycerate phosphatase and phosphoglycerate mutase activities are concurrently lost upon heating and treatment with reagents specific for histidyl, arginyl and lysyl residues. The two enzymes differ in their thermal stability and sensitivity to tetrathionate. Substrates and cofactors protect against inactivation, the protective effects varying with the modifying reagent. The synthase activity of both enzymes shows a nonhyperbolic pattern which fits to a second degree polynomial. The Km, Ki and optimum pH values are similar to those of the 2,3-bisphosphoglycerate synthase-phosphatase from erythrocytes and the hybrid enzyme from skeletal muscle. The synthase activity is inhibited by inorganic phosphate and it is stimulated by glycolyate 2-P.  相似文献   

8.
The levels of the enzymes involved in the metabolism of glycerate-2,3-P2 (phosphoglycerate mutase, bisphosphoglycerate synthase-phosphatase and bisphosphoglycerate phosphatase) in cat and in pig tissues are different. The main difference is the low level of bisphosphoglycerate synthase-phosphatase in cat tissues. As a consequence, in contrast with pig erythrocytes, in cat erythrocytes, both the synthesis and the breakdown of glycerate-2,3-P2 are mainly controlled by phosphoglycerate mutase.  相似文献   

9.
Bisphosphoglycerate mutase is an erythrocyte-specific enzyme catalyzing a series of intermolecular phosphoryl group transfer reactions. Its main function is to synthesize 2,3-bisphosphoglycerate, the allosteric effector of hemoglobin. In this paper, we directly observed real-time motion of the enzyme active site and the substrate during phosphoryl transfer. A series of high resolution crystal structures of human bisphosphoglycerate mutase co-crystallized with 2,3-bisphosphoglycerate, representing different time points in the phosphoryl transfer reaction, were solved. These structures not only clarify the argument concerning the substrate binding mode for this enzyme family but also depict the entire process of the key histidine phosphorylation as a "slow movie". It was observed that the enzyme conformation continuously changed during the different states of the reaction. These results provide direct evidence for an "in line" phosphoryl transfer mechanism, and the roles of some key residues in the phosphoryl transfer process are identified.  相似文献   

10.
The structure of a new crystal form of Saccharomyces cerevisiae phosphoglycerate mutase has been solved and refined to 2.12 A with working and free R-factors of 19.7 and 22.9 %, respectively. Higher-resolution data and greater non-crystallographic symmetry have produced a more accurate protein structure than previously. Prominent among the differences from the previous structure is the presence of two sulphate ions within each active site cleft. The separation of the sulphates suggests that they may occupy the same sites as phospho groups of the bisphosphate ligands of the enzyme. Plausible binding modes for 2,3-bisphosphoglycerate and 1, 3-bisphosphoglycerate are thereby suggested. These results support previous conclusions from mutant studies, highlight interesting new targets for mutagenesis and suggest a possible mechanism of enzyme phosphorylation.  相似文献   

11.
In pig skeletal muscle exist four enzymes with 2,3-bisphosphoglycerate phosphatase activity. Two of them (forms I-A and I-C) are multi-functional enzymes which, in addition to the phosphatase activity, possess 2,3-bisphosphoglycerate synthase and phosphoglycerate mutase activities. The other two enzyme forms (II-A and II-B) only show the phosphatase activity. The four enzymes differ in substrate specificity. Form I-C is highly specific for glycerate 2,3-P2; form I-A also hydrolyzes the monophosphoglycerates and forms II-A and II-B are specific for phosphoester bonds adjacent to a C-1 carboxylic group. The enzymes possess similar Km, Kcat and optimum pH value, but they are differently inhibited by the reaction products. They are also differently affected by glycolate-2-P (their main activator) and by other modifiers. Probably form I-A, which corresponds to M-type phosphoglycerate mutase, is the main enzyme implicated in the breakdown of glycerate 2,3-P2 in pig muscle.  相似文献   

12.
A system has been developed to allow the convenient production, expression and purification of site-directed mutants of the enzyme phosphoglycerate mutase from Saccharomyces cerevisiae. This enzyme is well characterised; both the amino acid sequence and crystal structure have been determined and a reaction mechanism has been proposed. However, the molecular basis for catalysis remains poorly understood, with only circumstantial evidence for the roles of most of the active site residues other than His8, which is phosphorylated during the reaction cycle. A vector/host expression system has been designed which allows recombinant forms of phosphoglycerate mutase to be efficiently expressed in yeast with no background wild-type activity. A simple one-column purification protocol typically yields 30 mg pure enzyme/1 l of culture. The active-site residue, His181, which is thought to be involved in proton transfer during the catalytic cycle, has been mutated to an alanine. The resultant mutant has been purified and characterised. Kinetic analysis shows a large decrease (1.6 x 10(4)) in the catalytic efficiency, and an 11-fold increase in the Km for the cofactor 2,3-bisphosphoglycerate. These observations are consistent with an integral role for His181 in the reaction mechanism of phosphoglycerate mutase, probably as a general acid or base.  相似文献   

13.
Kinetic analyses indicate that human erythrocyte phosphoglycerate mutase catalyzes the normal, reversible isomerization of D-glycerate-3-P and D-glycerate-2-P in the absence of added D-glycerate-2,3-P2. The reaction is impeded, however, by a potent inhibitor which occurs as a natural component of commericial D-glycerate-3-P. Inhibition may be overcome through substrate purification or by adding D-glycerate-2,3-P2 to the reaction medium containing the contaminant. In surmounting the inhibition, bisphosphoglycerate performs as a non-essential activator and not as a cofactor. The latter concept is corroborated by the observation that D-glycerate-2,3-P2 has absolutely no influence on mutase catalysis conducted in the presence of pure substrate. The data presented here and elsewhere, however, suggest that the red cell enzyme is readily phosphorylated by mono- as well as bisphosphoglycerate. Additional findings show that at concentrations in excess of 3mM, D-glycerate-3-P accelerates phosphoglycerate mutase catalysis in the absence of cofactor, suggesting that the mutase molecule possesses a normal catalytic site and an allosteric activator site.  相似文献   

14.
The primary sequence of maize 2,3-bisphosphoglycerate-independent phosphoglycerate mutase was deduced from cDNAs isolated from maize cDNA libraries by screening with specific antibodies to the cofactor-independent enzyme and from a maize genomic clone. The genomic clone provided the 5'-nucleotide sequence encoding the N-terminal amino acids which could not be obtained from the cDNA. Confirmation that the nucleotide sequence was for the cofactor-independent phosphoglycerate mutase was obtained by sequencing the peptides generated from cyanogen bromide cleavage of the purified protein. This is the first report of the amino acid sequence of a 2,3-bisphosphoglycerate cofactor-independent phosphoglycerate mutase, which consists of 559 amino acids and is twice the molecular size of the mammalian cofactor-dependent enzyme subunit. Analysis of the cofactor-independent phosphoglycerate mutase amino acid sequence revealed no identity with the cofactor-dependent mutase types. Northern blot analysis confirmed this difference since the maize cofactor-independent phosphoglycerate mutase cDNA did not hybridize with mRNA of the cofactor-dependent mutase. The lack of amino acid identity between cofactor-dependent and -independent enzymes is consistent with their different catalytic mechanisms and suggests that both enzymes are unrelated evolutionarily and arose from two independent ancestral genes. However, a constellation of residues which are involved in metal ion binding in various alkaline phosphatases is conserved in the maize cofactor-independent phosphoglycerate mutase, which suggests that the enzyme is a member of the alkaline phosphatase family of enzymes.  相似文献   

15.
The denaturation by guanidinium chloride of the monomeric phosphoglycerate mutase from Schizosaccharomyces pombe was studied. The loss in activity broadly parallels the changes in protein structure detected by fluorescence and c.d. Renaturation can be brought about by dilution of the denaturing agent. These processes were compared with those in the enzymes from baker's yeast and rabbit muscle, which are tetrameric and dimeric respectively. The effects of the cofactor 2,3-bisphosphoglycerate on the structure and stability of the S. pombe enzyme were also investigated.  相似文献   

16.
In contrast to the species with erythrocytes of high 2,3-bisphosphoglycerate content, in the sheep the concentration of 2,3-bisphosphoglycerate decreases during maturation of reticulocytes. The decrease can be explained by the drop of the phosphofructokinase/pyruvate kinase and 2,3-bisphosphoglycerate synthase/2,3-bisphosphoglycerate phosphatase activity ratios that result from the decline of phosphofructokinase, pyruvate kinase, phosphoglycerate mutase and the bifunctional enzyme 2,3-bisphosphoglycerate synthase/phosphatase. The concentrations of fructose 2,6-bisphosphate and aldohexose 1,6-bisphosphates also decrease during sheep reticulocyte maturation in parallel to the 6-phosphofructo 2-kinase and the glucose 1,6-bisphosphate synthase activities.  相似文献   

17.
Phosphoglycerate mutase and bisphosphoglycerate synthase (mutase) can both be phosphorylated by either glycerate-1,3-P2 or glycerate-2,3-P2 to form phosphohistidine enzymes. The present study uses a rapid quench procedure to determine if, for each enzyme, the formation of the phosphorylated enzyme and phosphate transfer from the enzyme can occur at rates consistent with the overall reactions. With bisphosphoglycerate synthase from horse red blood cells (glycerate-1,3-P2 leads to glycerate-2,3-P2) at pH 7.5, 25 degrees, phosphorylation of the enzyme appears rate-limiting, k = 13.5 s-1, compared with kcat = 12.5 s-1 for the overall synthase rate. Phosphoryl transfer from the enzyme to phosphoglycerate occurs at 38 s-1 at 4 degrees and was too fast to measure at 25 degrees. With chicken muscle phosphoglycerate mutase the half-times were too short to measure under optimal conditions. The rate of enzyme phosphorylation by glycerate-2,3-P2 at pH 5.5, 4 degrees, could account for the overall reaction rate of 170 s-1. The rate of phosphoryl transfer from the enzyme to glycerate-3-P was too rapid to measure under the same conditions. It is concluded that the phosphorylated enzymes have kinetic properties consistent with their participation as intermediates in the reactions catalyzed by these enzymes.  相似文献   

18.
The steady-state kinetics of 1,3-bisphosphoglycerate formation through the action of phosphoglycerate kinase on 3-phosphoglycerate and ATP have been examined. The results show that initial velocities determined by the standard method of coupling bisphosphoglycerate production to NADH reduction in the presence of glyceraldehyde-3-phosphate dehydrogenase do not differ significantly from those determined in the absence of the latter enzyme. This observation invalidates the proposal that bisphosphoglycerate dissociation from phosphoglycerate kinase is much too slow to account for the high rates of phosphoglycerate turnover observed in the coupled two-enzyme system. The capacity for rapid bisphosphoglycerate production and release is an intrinsic catalytic property of phosphoglycerate kinase that does not require the presence of other enzymes or the involvement of a mechanism of channelized (non-diffusional) transfer of bisphosphoglycerate from the producing enzyme to the consuming one.  相似文献   

19.
Vanadate causes a rapid breakdown of 2,3-bisphosphoglycerate in intact erythrocytes. This metabolite is nearly stoichiometrically transformed into pyruvate, which changes the cell redox state and enhances the glycolytic flux. The results show that the vanadate effect on 2,3-bisphosphoglycerate, also evident in hemolysates, is attributable to the stimulation of a phosphatase activity of the phosphoglycerate mutase. In agreement with others (J. Carreras, F. Climent, R. Bartrons and G. Pons (1982) Biochim. Biophys. Acta705, 238–242), vanadate is thought to destabilize the phosphoryl form of this enzyme which shows competitive inhibition between the ion and 2,3-bisphosphoglycerate in the mutase reaction. A competitive inhibition between vanadate and glucose 1,6-bisphosphate is also found for phosphoglucomutase, without evidence for phosphatase activity toward the bisphosphate cofactor.  相似文献   

20.
In the rabbit and in the rat, which possess erythrocytes with high concentration of 2,3-bisphosphoglycerate, the 2,3-bisphosphoglycerate synthase activity increases more than two fold during reticulocyte maturation. Isolation of the enzymes with 2,3-bisphosphoglycerate synthase activity present in extracts of reticulocytes and mature erytrocytes by ion exchange fast liquid chromatography shows that the increase in the synthase activity is due to the accumulation of the bifunctional enzyme 2,3-bisphosphoglycerate synthase/phosphatase (EC 2.7.5.4/EC 3.1.3.13) which represents more than 80% of the synthase activity of the cell extracts. During reticulocyte maturation phosphoglycerate mutase (EC 5.4.2.1), which makes a small contribution to the 2,3-bisphosphoglycerate synthase activity in the erythroid cells, decreases in the rabbit and remains constant in the rat.  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