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.
The lysine-183 residues of yeast glyceraldehyde 3-phosphate dehydrogenase, in contrast to the cysteine-149 residues, react independently with acylating and alkylating agents. Modification of all four residues is required to inactivate the enzyme in spite of the fact that this residue is apparently in the neighborhood of the cysteine-149 involved in half-of-the-sites activity. The modification of the lysine-183 residue, however, influences the half-of-the-sites effect since alkylation of the cysteine-149 residues of the enzyme whose lysine-183 residues are acetylated follows a linear pattern with each subunit acting independently. Four lysine residues outside the active site can be modified with fluorodinitrobenzene, causing 80% loss in enzyme activity. Once again each subunit acts independently. This same residue can also be modified by a fluorescein label which can serve as a reporter group for binding and conformational changes occurring at the active site. The results add support for the functional symmetry of the apo-enzyme and demonstrate how the co-operativity between subunits can be altered by amino acid modification.  相似文献   

2.
Glyceraldehyde 3-phosphate dehydrogenase (EC 1.2.1.12) from the extremely halophilic archaebacterium Haloarcula vallismortis has been purified in a four step procedure to electrophoretic homogeneity. The enzyme is a tetramer with a relative molecular mass of 160000. It is strictly NAD+-dependent and exhibits its highest activity in 2 mol/l KCl at 45°C. Amino acid analysis and isoelectric focusing indicate an excess of acidic amino acids. Two parts of the primary sequence are reported. These peptides have been compared with glyceraldehyde 3-phosphate dehydrogenases from other archaebacteria, eubacteria and eucaryotes. The peptides show a high grade of similarity to glyceraldehyde 3-phosphate dehydrogenase from eucaryotes.Abbreviations BCA bicinchoninic acid - CTAB cetyltrimethyl ammonium bromide - DTE dithioerythritol - DTT dithiothreitol - GAP glyccraldehyde 3-phosphate - GAPDH glyceraldehyde 3-phosphate dehydrogenase  相似文献   

3.
Pyrene maleimide is shown to be a 'half of the sites' reagent for glutamate dehydrogenase and for glyceraldehyde-3-phosphate dehydrogenase. The modified residues are identified as cysteine-115 for glutamate dehydrogenase and cysteine-149 for glyceraldehyde-3-phosphate dehydrogenase. The two enzymes react differently with pyrene maleimide. Whereas the hydrophobic environment of cysteine-115 directs the modification of glutamate dehydrogenase, the high reactivity of cysteine-149 determines the specific modification of glyceraldehyde-3-phosphate dehydrogenase. Glutamate dehydrogenase activity is unaltered by the modification: glyceraldehyde-3-phosphate dehydrogenase activity in inhibited.  相似文献   

4.
Glyceraldehyde 3-phosphate dehydrogenase and phosphoribulokinase exist as stable enzymes and as part of a complex in Chlamydomonas reinhardtii. We show here that phosphoribulokinase exerts an imprinting on glyceraldehyde 3-phosphate dehydrogenase, which affects its catalysis by decreasing the energy barrier of the reactions with NADH or NADPH by 3.8 +/- 0.5 and 1.3 +/- 0.3 kJ.mol(-1). Phosphoribulokinase and glyceraldehyde 3-phosphate dehydrogenase within the complex are regulated by NADP(H) but not by NAD(H). The activities of the metastable phosphoribulokinase and glyceraldehyde 3-phosphate dehydrogenase released from the complex preincubated with NADP(H) are different from those of the metastable enzymes released from the untreated complex. NADP(H) increases phosphoribulokinase and NADPH-glyceraldehyde 3-phosphate dehydrogenase activities with a (~)K(0.5 (NADP)) of 0.68 +/- 0.16 mm and a (~)K(0.5 (NADPH)) of 2.93 +/- 0.87 mm and decreases NADH-dependent activity. 1 mm NADP increases the energy barrier of the NADH-glyceraldehyde 3-phosphate dehydrogenase-dependent reaction by 1.8 +/- 0.2 kJ.mol(-1) and decreases that of the reactions catalyzed by phosphoribulokinase and NADPH-glyceraldehyde 3-phosphate dehydrogenase by 3 +/- 0.2 and 1.2 +/- 0.3 kJ.mol(-1), respectively. These cofactors have no effect on the independent stable enzymes. Therefore, protein-protein interactions may give rise to new regulatory properties.  相似文献   

5.
Glyceraldehyde 3-phosphate dehydrogenase exhibits half-site reactivity, the structural origin of which is obscure. Thermal inactivation kinetics, employed here as a probe for site-site heterogeneity in solution, show that green gram glyceraldehyde 3-phosphate dehydrogenase (in the absence and presence of phosphate and NAD+) loses activity in two distinct phases, each of which accounts for half of the initial activity. In the presence of substrate, glyceraldehyde 3-phosphate the relative amplitude of the slow phase increases, and at 0.06 mM glyceraldehyde 3-phosphate the time-course of inactivation corresponds to a single exponential decay. The data are consistent with a suggestion that glyceraldehyde 3-phosphate dehydrogenase may exist in two interconvertible conformations of different symmetry characteristics (C2 in equilibrium D2). The lower symmetry conformation (C2) predominates in the apoenzyme and in the presence of phosphate and NAD+. The higher symmetry conformation (D2) is stabilised by glyceraldehyde 3-phosphate.  相似文献   

6.
The interaction of glyceraldehyde 3-phosphate dehydrogenase with microtubules has been studied by measurement of the amount of enzyme which co-assembles with in vitro reconstituted microtubules. The binding of glyceraldehyde 3-phosphate dehydrogenase to microtubules is a saturable process; the maximum binding capacity is about 0.1 mole of enzyme bound per mole of assembled tubulin. Half saturation of microtubule binding sites is obtained at a concentration of glyceraldehyde 3-phosphate dehydrogenase of about 0.5 µM Glyceraldehyde 3-phosphate dehydrogenase (between 0.1 and 2 µM) induces a concentration-dependent increase a) in the turbidity of the microtubule suspension without alteration of the net amount of polymer formed and b) in the amount of microtubule protein polymers after cold microtubule disassembly. There is a linear relationship between the intensity of the glyceraldehyde 3-phosphate dehydrogenase-induced effects and the amount of microtubule-bound enzyme. The specificity of the association of glyceraldehyde 3-phosphate dehydrogenase to microtubules has been documented by copolymerization experiments. Assembly-disassembly cycles of purified microtubules in the presence of a crude liver soluble fraction results in the selective extraction of a protein with an apparent molecular weight of 35 000 identified as the monomer of glyceraldehyde 3-phosphate dehydrogenase by peptide mapping and immunoblotting.In conclusion, microtubules possess a limited number of binding sites for glyceraldehyde 3-phosphate dehydrogenase. The binding of the glycolytic enzyme to microtubules shows a considerable specificity and is associated with alterations of assembly and disassembly characteristics of microtubules.Abbreviations Mes 2(N-morpholinoethane) sulfonic acid - EGTA ethylene glycol bis (-aminoethyl-ester)N,N,N,N tetraacetic acid - EDTA thylene diamine tetraacetic acid  相似文献   

7.
A procedure is presented for the large-scale purification of rabbit muscle glyceraldehyde 3-phosphate dehydrogenase (EC 1.2.1.12). Sephadex G-100 chromatography was more effective than repeated recrystallizations for removing heme impurities, hence these proteins appear to cocrystallize. The column purified enzyme has full enzymatic activity according to dehydrogenase, esterase, and acetyl phosphatase assays.  相似文献   

8.
Covalent modification of two of the four cysteine-149 residues of yeast glyceraldehyde 3-phosphate dehydrogenase, at pH 8.5, decreases the reactivity of the remaining two cysteine-149 residues and essentially inactivates the protein. The structure of the modifying reagent has only a secondary influence on this half-of-the-sites effect. Reactivity studies, together with the existing X-ray and sequence studies, suggest that the four active sites are initially functionally identical both in activity and in cysteine reactivity. The half-of-the-sites effect therefore arises in part or in whole from ligand-induced negatively co-operative conformational changes. A detailed kinetic study with iodoacetamide gives relative values of two rapidly reacting groups, a third more slowly reacting, and a fourth very slowly reacting group. These data, added to the existing data on cytidine triphosphate synthetase and alkaline phosphatase, suggest that the half-of-the-sites phenomena in many enzymes may be explained by ligand-induced negative co-operativity triggered by binding or covalent bond formation or both.  相似文献   

9.
1. The amino acid sequence of D-glyceraldehyde-3-phosphate dehydrogenase from the extreme thermophile Thermus aquaticus has been elucidated. 2. The polypeptide contains 332 amino acids and its sequence is 70% identical with that of the enzyme from the moderate thermophile Bacillus stearothermophilus. 3. In contrast to less thermostable forms of the enzymes from B. stearothermophilus, pig, lobster and yeast, the T. aquaticus enzyme has only one cysteine residue, namely cysteine-149 which is required for catalysis.  相似文献   

10.
Bischloromethylpentanedione, bischloromethylhexanedione, bischloromethyloctanedione and bischloromethyldecanedione were synthesized from their corresponding dicarboxylic acids via the bis-acyl chloride and the bisdiazomethylketone derivatives. These compounds proved to be highly specific cross-linking reagents for rabbit skeletal-muscle glyceraldehyde 3-phosphate dehydrogenase. Incubation of the enzyme with cross-linking reagents resulted in both a time- and concentration-dependent formation of covalently linked oligomeric structures. The major cross-linked product detected by sodium dodecyl sulphate/polyacrylamide-gel electrophoresis was the dimer (mol. wt. 72000). Sepharose 6B chromatography of the cross-linked enzyme showed that it still existed as the tetramer. Cross-linking was dependent on the native structure of the enzyme, since it was abolished on denaturation of the enzyme. The actual covalently linked product depends on the conditions of modification and the chain length of the reagent. The maximum yield of dimer (70-80%) was obtained with bischloromethylhexanedione, and the yield decreased with either shorter- or longer-chain compounds. The calculated distance between the two reactive points in bischloromethylhexanedione is 1.21-1.45nm. Bischloromethylhexanedione modified at least two thiol groups per monomer. Modification of the active-site thiol, cysteine-149, was not essential for cross-linking, since glyceraldehyde 3-phosphate dehydrogenase carboxymethylated on cysteine-149 still reacted to form the dimer. The rate of chemical cross-linking was markedly decreased by increasing the NAD(+) occupancy of the enzyme active sites. These experiments are discussed in terms of the asymmetry of the enzyme structure in solution.  相似文献   

11.
Glyceraldehyde 3-phosphate dehydrogenase, a tetramer of 140,000 Da, interacts with in vitro reconstituted microtubules. It results in a partial inhibition of the activity of the microtubule-bound enzyme. After cold depolymerization of the microtubule-glyceraldehyde 3-phosphate dehydrogenase complexes, a fraction of the enzyme is recovered in an active form in the disassembly supernatant; the other fraction devoid of activity, identified by polyacrylamide gel electrophoresis, remains associated with the undepolymerizable microtubule protein pellet. The inactivation of the microtubule-bound enzyme is related to the concentration of microtubule protein. Higher the concentration of microtubule protein, lower the fraction of inactivated enzyme; consequently, glyceraldehyde 3-phosphate dehydrogenase is able to copolymerize quantitatively with microtubule protein through one assembly-disassembly cycle, provided that the concentration of microtubule protein is high. Monomeric glyceraldehyde 3-phosphate dehydrogenase (molecular weight: 35,000) devoid of enzyme activity, prepared by reversible dissociation of the tetrameric enzyme, also binds to microtubules and is quantitatively recovered in the undepolymerizable microtubule protein fraction after cold treatment. These results indicate that interacting with microtubules, glyceraldehyde 3-phosphate dehydrogenase partly dissociates into inactive monomers, this process is regulated by the concentration of assembled microtubule protein, and active and inactive glyceraldehyde 3-phosphate dehydrogenase bound to microtubules have different fate at the step of microtubule disassembly. These data suggest that an association of glyceraldehyde 3-phosphate dehydrogenase to microtubules could play a role in modulating the activity of the glycolytic enzyme in intact cells.  相似文献   

12.
Rat muscle glyceraldehyde-3-phosphate dehydrogenase was reacted with two reagents aimed at the highly reactive cysteine-149 residue in the active site of the enzyme. The enzyme was rapidly inactivated by iodine monochloride. Complete inactivation occured when approx. 6 mol ICl were added per mol enzyme, indicating that reactions which compete with the reagent's interaction with cysteine-149 take place. Iodine was also found to inactivate the enzyme rapidly and effectively, and, when not in excess, this reagent interacted specifically with cysteine-149. The fraction of original enzymatic activity which could be restored by 2-mercaptoethanol in enzyme samples inhibited by 4.2 mol I2/mol enzyme, decreased with time to a limiting value of 0.6 reached after approx. 15 min. The enzyme thus treated showed a remarkable similarity to enzyme samples purified from old rats, both in its activity and in NAD+ binding patterns under various conditions. It is concluded that the structural modifications induced in the modified enzyme resemble the age-related modifications in native ‘old’ enzyme. These results demonstrate that the origin of the age-related effects in glyceraldehyde-3-phosphate dehydrogenase is in subtle, post-synthetic structural changes. The inactivation reactions described above require a non-reducing environment for the enzyme. Whether such conditions do exist in cells of old animals is the subject of future studies.  相似文献   

13.
Perdeuterated spin label (DSL) analogs of NAD+, with the spin label attached at either the C8 or N6 position of the adenine ring, have been employed in an EPR investigation of models for negative cooperativity binding to tetrameric glyceraldehyde-3-phosphate dehydrogenase and conformational changes of the DSL-NAD+-enzyme complex during the catalytic reaction. C8-DSL-NAD+ and N6-DSL-NAD+ showed 80 and 45% of the activity of the native NAD+, respectively. Therefore, these spin-labeled compounds are very efficacious for investigations of the motional dynamics and catalytic mechanism of this dehydrogenase. Perdeuterated spin labels enhanced spectral sensitivity and resolution thereby enabling the simultaneous detection of spin-labeled NAD+ in three conditions: (1) DSL-NAD+ freely tumbling in the presence of, but not bound to, glyceraldehyde-3-phosphate dehydrogenase, (2) DSL-NAD+ tightly bound to enzyme subunits remote (58 A) from other NAD+ binding sites, and (3) DSL-NAD+ bound to adjacent monomers and exhibiting electron dipolar interactions (8-9 A or 12-13 A, depending on the analog). Determinations of relative amounts of DSL-NAD+ in these three environments and measurements of the binding constants, K1-K4, permitted characterization of the mathematical model describing the negative cooperativity in the binding of four NAD+ to glyceraldehyde-3-phosphate dehydrogenase. For enzyme crystallized from rabbit muscle, EPR results were found to be consistent with the ligand-induced sequential model and inconsistent with the pre-existing asymmetry models. The electron dipolar interaction observed between spin labels bound to two adjacent glyceraldehyde-3-phosphate dehydrogenase monomers (8-9 or 12-13 A) related by the R-axis provided a sensitive probe of conformational changes of the enzyme-DSL-NAD+ complex. When glyceraldehyde-3-phosphate was covalently bound to the active site cysteine-149, an increase in electron dipolar interaction was observed. This increase was consistent with a closer approximation of spin labels produced by steric interactions between the phosphoglyceryl residue and DSL-NAD+. Coenzyme reduction (DSL-NADH) or inactivation of the dehydrogenase by carboxymethylation of the active site cysteine-149 did not produce changes in the dipolar interactions or spatial separation of the spin labels attached to the adenine moiety of the NAD+. However, coenzyme reduction or carboxymethylation did alter the stoichiometry of binding and caused the release of approximately one loosely bound DSL-NAD+ from the enzyme. These findings suggest that ionic charge interactions are important in coenzyme binding at the active site.  相似文献   

14.
The catalytically essential amino acid, histidine 176, in the active site of Escherichia coli glyceraldehyde-3-phosphate dehydrogenase (GAPDH) has been replaced with an asparagine residue by site-directed mutagenesis. The role of histidine 176 as a chemical activator, enhancing the reactivity of the thiol group of cysteine 149, has been demonstrated, with iodoacetamide as a probe. The esterolytic properties of GAPDH, illustrated by the hydrolysis of p-nitrophenyl acetate, have been also studied. The kinetic results favor a role for histidine 176 not only as a chemical activator of cysteine 149 but also as a hydrogen donor facilitating the formation of tetrahedral intermediates. These results support the hypothesis that histidine 176 plays a similar role during the oxidative phosphorylation of glyceraldehyde 3-phosphate.  相似文献   

15.
A procedure is described for the purification of three glyceraldehyde phosphate dehydrogenases from a batch of beet leaves. Glyceraldehyde 3-phosphate:NADP+ reductase, nonphosphorylating (EC 1.2.1.9) has been purified over 1500-fold. The Mr of this enzyme is 190,000 and its subunits have an Mr of 53,000, suggesting a tetramer as the active form. Its pI is 6.0. Cytosolic glyceraldehyde 3-phosphate dehydrogenase, NAD dependent (EC 1.2.1.12), has an Mr of 145,000 and subunits of Mr 37,000. It is dissociated to inactive dimers by ATP, whereas NAD+ in the presence of reductant promotes its reactivation. The amino acid composition is related to glyceraldehyde 3-phosphate dehydrogenases from animal sources and is most similar to pea seed glyceraldehyde 3-phosphate dehydrogenase. The enzyme exhibits a range of pI values from 5 to 7, but a second electrofocusing in the presence of dithioerythritol results in a single main form with pI 5.33, consistent with the behavior in polyacrylamide and cellulose acetate gel electrophoresis. Chloroplast NAD(P)-glyceraldehyde 3-phosphate dehydrogenase (EC 1.2.1.13) has been obtained from beet, pea, Ranunculus, Arum, and maize leaves. The stable form is an oligomer of about 800,000 Mr (±10%), while a minor, possibly damaged fraction elutes as a retarded peak from agarose columns. The Mr 800,000 form is reversibly dissociated to protomers of Mr 160,000 by NADP+, with increase of apparent NADP-dependent activity. Two subunits are present in similar amounts in all association states and after all treatments: α with Mr 36,000, and β with Mr 41,000. The form found in density gradient ultracentrifugation has an Mr of 390,000. Isoelectric points of the various forms lie between pH 4.1 and 4.7 for all species, with a main peak usually at pI 4.45. The amino acid composition of beet chloroplast glyceraldehyde phosphate dehydrogenase is not closely related to that of beet leaf NAD-dependent glyceraldehyde 3-phosphate dehydrogenase.  相似文献   

16.
PLP synthase (PLPS) is a remarkable single-enzyme biosynthetic pathway that produces pyridoxal 5′-phosphate (PLP) from glutamine, ribose 5-phosphate, and glyceraldehyde 3-phosphate. The intact enzyme includes 12 synthase and 12 glutaminase subunits. PLP synthesis occurs in the synthase active site by a complicated mechanism involving at least two covalent intermediates at a catalytic lysine. The first intermediate forms with ribose 5-phosphate. The glutaminase subunit is a glutamine amidotransferase that hydrolyzes glutamine and channels ammonia to the synthase active site. Ammonia attack on the first covalent intermediate forms the second intermediate. Glyceraldehyde 3-phosphate reacts with the second intermediate to form PLP. To investigate the mechanism of the synthase subunit, crystal structures were obtained for three intermediate states of the Geobacillus stearothermophilus intact PLPS or its synthase subunit. The structures capture the synthase active site at three distinct steps in its complicated catalytic cycle, provide insights into the elusive mechanism, and illustrate the coordinated motions within the synthase subunit that separate the catalytic states. In the intact PLPS with a Michaelis-like intermediate in the glutaminase active site, the first covalent intermediate of the synthase is fully sequestered within the enzyme by the ordering of a generally disordered 20-residue C-terminal tail. Following addition of ammonia, the synthase active site opens and admits the Lys-149 side chain, which participates in formation of the second intermediate and PLP. Roles are identified for conserved Asp-24 in the formation of the first intermediate and for conserved Arg-147 in the conversion of the first to the second intermediate.  相似文献   

17.
C Corbier  F Della Seta  G Branlant 《Biochemistry》1992,31(49):12532-12535
NAD(P) aldehyde dehydrogenases (EC 1.2.1.3) are a family of enzymes that oxidize a wide variety of aldehydes into acid or activated acid compounds. Using site-directed mutagenesis, the essential nucleophilic Cys 149 in the NAD-dependent phosphorylating glyceraldehyde-3-phosphate dehydrogenase from Escherichia coli has been replaced by alanine. Not unexpectedly, the resulting mutant no longer shows any oxidoreduction phosphorylating activity. The same mutation, however, endows the enzyme with a novel oxidoreduction nonphosphorylating activity, converting glyceraldehyde 3-phosphate into 3-phosphoglycerate. Our study further provides evidence for an alternative mechanism in which the true substrate is the gem-diol entity instead of the aldehyde form. This implies that no acylenzyme intermediate is formed during the catalytic event. Therefore, the mutant C149A is a new enzyme which catalyzes a distinct reaction with a chemical mechanism different from that of its parent phosphorylating glyceraldehyde-3-phosphate dehydrogenase. This finding demonstrates the possibility of an alternative route for the chemical reaction catalyzed by classical nonphosphorylating aldehyde dehydrogenases.  相似文献   

18.
Human lung carcinoma cells (A549) were oxidatively stressed with mildly-toxic or non-toxic amounts of hydrogen peroxide (H2O2, 0.1 mM to 120 mM) for 5 min. Hydrogen peroxide exposure resulted in a dose dependent inhibition of binding (pH 7) of the thiol reagent iodoacetic acid (IAA) to a 38 kDa cell protein. Incubation of cells in saline for 60 min following H2O2 removal restored the ability of IAA to bind to the protein. Treatment with 20 mM dithiothreitol or 2 M urea also restored IAA binding, but 10% Triton X102 or 1 mM Brij 58 had no effect. Increasing to pH 11 during the IAA binding also increased thiol availability. Glyceraldehyde 3-phosphate dehydrogenase (EC 1.2.1.12) has been identified as the protein undergoing thiol/disulfide redox status and enzymic activity changes.  相似文献   

19.
Crystalline d-glyceraldehyde 3-phosphate dehydrogenase from lobster tail contains 4 moles of NAD(+) bound and reacts specifically with 4 moles of iodoacetic acid/mole of tetramer. The essential thiol group of d-glyceraldehyde 3-phosphate dehydrogenase appears to react with iodoacetic acid with a rate constant for the overall process that is independent of the extent of carboxymethylation. The d-glyceraldehyde 3-phosphate dehydrogenase-NAD(+) absorption band has a variable molar extinction coefficient in the presence of phosphate that may be correlated with a proton dissociation of pK 6.86. The binding of NAD(+) to d-glyceraldehyde 3-phosphate dehydrogenase weakens as alkylating agents react with the enzyme, and NAD(+) promotes the reactivity of the essential thiol group. It is suggested that, on binding to d-glyceraldehyde 3-phosphate dehydrogenase, NAD(+) lowers the pK of the essential thiol group, resulting in a catalytic role of NAD(+) in the reaction catalysed by d-glyceraldehyde 3-phosphate dehydrogenase. If this theory is correct, then it is likely that a proton will be liberated during the phosphorolysis of the acyl-enzyme rather than in the redox step.  相似文献   

20.
Glyceraldehyde induces changes in the flux of glucose oxidised through the hexose monophosphate pathway, the concentrations of intermediates in the Embden-Meyerhoff pathway, the oxidative status of haemoglobin and levels of reduced and oxidised pyridine nucleotides and glutathione in red cells. Glyceraldehyde autoxidises in the cellular incubations, consuming oxygen and producing glyoxalase I- and II-reactive materials. Major fates of glyceraldehyde in red cells appear to be: (i) adduct formation with reduced glutathione and cellular protein; (ii) autoxidation and reaction with oxyhaemoglobin and pyridine nucleotides, and (iii) phosphorylation of D-glyceraldehyde and entry into the glycolytic pathway as glyceraldehyde 3-phosphate. The production of glycerol from glyceraldehyde by red cell L-hexonate dehydrogenase appears not to be a major reaction of glyceraldehyde in red cells. These results indicate that high concentrations of glyceraldehyde (1-50 mM) may induce oxidative stress in red cells by virtue of the spontaneous autoxidation of glyceraldehyde, forming hydrogen peroxide and alpha-ketoaldehydes (glyoxalase substrates). The implications of glyceraldehyde-induced oxidative stress for the in vitro anti-sickling effect of DL-glyceraldehyde and for the polyol pathway metabolism of glyceraldehyde are discussed.  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