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 tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle binds NAD+ and some of its analogues in a negatively cooperative manner, whereas other NAD+ analogues bind non-cooperatively to this enzyme. Subsequent to alkylation of a fraction of the active sites of the enzyme with the fluorescent SH reagent N-iodoacetyl-N'-(5-sulfo-1-naphthyl)-ethylenediamine, it was found that the alkylated sites bind NAD+ and NAD+ analogues with a markedly reduced affinity as compared with non-alkylated sites. It was therefore feasible to measure the fluorescence and the circular polarization of the luminescence of the enzyme-bound alkyl groups as a function of binding of NAD+ and of NAD+ analogues to the non-alkylated sites. The changes observed indicate that ligand binding to the non-alkylated sites induces changes in the fluorescence properties of the alkyl groups bound to neighbouring subunits, most likely through the protein moiety. The nature of these changes appears to depend on the structure of the coenzyme analogue. The binding of the non-cooperative binders acetyl-pyridine--adenine dinucleotide, ATP and ADP-ribose induce different conformational changes in the neighbouring vacant subunit, as monitored by the spectroscopic properties of the bound alkyl group. These results in conjunction with other data support the view that the negative cooperativity in NAD+ binding to glyceraldehyde-3-phosphate dehydrogenase results from ligand-induced conformational changes. Furthermore, these results further support the view that subtle structural changes in the coenzyme molecule determine the nature of the conformational changes induced within the enzyme tetramer.  相似文献   

2.
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   

3.
This work reports on the interaction of the fluorescent nicotinamide 1,N6-ethenoadenine dinucleotide (epsilonNAD+) with horse liver alcohol dehydrogenase, octopine dehydrogenase, and glyceraldehyde-3-phosphate dehydrogenase from different sources (yeast, lobster muscle, and rabbit muscle). The coenzyme fluorescence is enhanced by a factor of 10-13 in all systems investigated. It is shown that this enhancement cannot be due to changes in the polarity of the environment upon binding, and that it must be rather ascribed to structural properties of the bound coenzyme. Although dynamic factors could also be important for inducing changes in the quantum yield of epsilonNAD+ fluorescence, the close similarity of the fluorescence enhancement factor in all cases investigated indicates that the conformation of bound coenzyme is rather invariant in the different enzyme systems and overwhelmingly shifted toward an open form. Dissociation constants for epsilonNAD+-dehydrogenases complexes can be determined by monitoring the coenzyme fluorescence enhancement or the protein fluorescence quenching. In the case of yeast glyceraldehyde-3-phosphate dehydrogenase at pH 7.0 and t = 20 degrees the binding plots obtained by the two methods are coincident, and show no cooperativity. The affinity of epsilonNAD+ is generally lower than that of NAD+, although epsilonNAD+ maintains most of the binding characteristics of NAD+. For example, it forms a tight complex with horse liver alcohol dehydrogenase and pyrazole, and with octopine dehydrogenase saturated by L-arginine and pyruvate. One major difference in the binding behavior of NAD+ and epsilonNAD+ seems to be present in the muscle glyceraldehyde-3-phosphate dehydrogenase. In fact, no difference was found for epsilon NAD+ between the affinities of the third and fourth binding sites. The results and implications of this work are compared with those obtained recently by other authors.  相似文献   

4.
The binding of NAD+ to glyceraldehyde 3-phosphate dehydrogenase (EC 1.2.1.12) from Bacillus stearothermophilus has been studied by measurement of protein fluorescence quenching. Slight negative co-operativity was observed in the binding of the third and fourth coenzyme molecules to the tetrameric enzyme. The first two coenzyme molecules were tightly bound. In this respect the enzyme resembles that from sturgeon muscle rather than that from yeast.  相似文献   

5.
Directed mutagenesis has been used to study the nicotinamide subsite of the glycolytic NAD-dependent glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Residue Asn313 is involved together with the carboxyamide moiety of the nicotinamide ring in a complex network of hydrogen bonding interactions which fix the position of the pyridinium ring of NAD to which hydride transfer occurs at the C-4 position in the catalytic reaction. The asparagine side-chain has been replaced by that of the Thr and Ala residues and results in mutants with very similar properties. Both mutants show much weaker binding of NAD and lower catalytic efficiency. The mutant Asn313----Thr still exhibits strict B-stereospecificity in hydride transfer and retains the property of negative co-operativity in NAD binding. These experiments strongly suggest that the mutant enzyme undergoes the apo----holo sub-unit structural transition associated with coenzyme binding but that the nicotinamide ring is no longer as rigidly held in its pocket as in the wild type enzyme. The results shed light on the details of the molecular interactions which are responsible for negative co-operativity in this enzyme.  相似文献   

6.
In the course of studying mammalian erythrocytes we noted prominent differences in the red cells of the rat. Analysis of ghosts by sodium dodecyl sulfate (SDS) polyacrylamide gel electrophoresis showed that membranes of rat red cells were devoid of band 6 or the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12). Direct measurements of this enzyme showed that glyceraldehyde-3-phosphate dehydrogenase activity in rat erythrocytes was about 25% of that in human cells; all of the glyceraldehyde-3-phosphate dehydrogenase activity in rat erythrocytes was within the cytoplasm and none was membrane bound; and in the human red cell, about 1/3 of the enzyme activity was within the cytoplasm and 2/3 membrane bound. The release of glyceraldehyde-3-phosphate dehydrogenase from fresh rat erythrocytes immediately following saponin lysis was also determined using the rapid filtration technique recently described. The extrapolated zero-time intercepts of these reactions confirmed that, in the rat erythrocyte, none of the cellular glyceraldehyde-3-phosphate dehydrogenase was membrane bound. Failure of rat glyceraldehyde-3-phosphate dehydrogenase to bind to the membranes of the intact rat erythrocyte seems to be due to cytoplasmic metabolites which interact with the enzyme and render it incapable of binding to the membrane.  相似文献   

7.
The stabilizing effect of the coenzyme (NAD) on the structure of glyceraldehyde-3-phosphate dehydrogenase from lamprey and porcine muscles with respect to proteolysis and heat denaturation was studied. The process of heat denaturation was followed by the changes in specific activity of the enzymes; that of proteolysis--by the changes in specific activity and circular dichroism. It was shown that in both cases NAD at saturating concentration exerts a far weaker stabilizing effect on the structure of glyceraldehyde-3-phosphate dehydrogenase from lamprey muscle than on that of the porcine muscle enzyme. The coensyme-dependent stabilization of lamprey muscle glyceraldehyde-3-phosphate dehydrogenase does not differ from that of mammalian muscle enzyme. Possible interrelationship between the phenomenon observed and the molecular mechanism of thermal adaptation in the cold-blooded animals is discussed.  相似文献   

8.
There is no evidence, at pH 9.4, of negative cooperativity in the binding of NAD+ or NADH to rabbit muscle glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phorphorylating), EC 1.2.1.12) nor in the binding of acetyl pyridine adenine dinucleotide at pH 7.6 and ph 9.4. The binding of NAD+ to carboxymethylated enzyme at pH 7.6 and pH 9.4 also occurs without cooperativity. The possible implications of these findings for the involvement of ionising groups in the enzyme in the subunit interactions responsible for negative cooperativity, previously reported for coenzyme binding at pH 7.4--8.6, are discussed.  相似文献   

9.
The secondary structure of glycerol-3-phosphate dehydrogenase was predicted from its amino acid sequence. The pattern of helices and sheets within the first half of the polypeptide as well as specific marker residues were consistent with the properties of the NAD binding domain in other dehydrogenases. The second half of the sequence shows similarities with the catalytic domain of glyceraldehyde-3-phosphate dehydrogenase. The resulting two-domain structure of glycerol-3-phosphate dehydrogenase allows the correct environment for the B specificity of the nicotinamide ring and the L-glycerol 3-phosphate substrate.  相似文献   

10.
S Chen  T D Lee  K Legesse  J E Shively 《Biochemistry》1986,25(19):5391-5395
We have identified the site labeled by arylazido-beta-alanyl-NAD+ (A3'-O-(3-[N-(4-azido-2-nitrophenyl)amino]propionyl)NAD+) in rabbit muscle glyceraldehyde-3-phosphate dehydrogenase by microsequencing and fast atom bombardment mass spectrometry. This NAD+ photoaffinity analogue has been previously demonstrated to modify glyceraldehyde-3-phosphate dehydrogenase in a very specific manner and probably at the active site of the enzyme [Chen, S., Davis, H., Vierra, J. R., & Guillory, R. J. (1984) Biochem. Biophys. Stud. Proteins Nucleic Acids, Proc. Int. Symp., 3rd, 407-425]. The label is associated exclusively with a tryptic peptide that has the sequence Ile-Val-Ser-Asn-Ala-Ser-Cys-Thr-Thr-Asn. In comparison to the amino acid sequence of glyceraldehyde-3-phosphate dehydrogenase from other species, this peptide is in a highly conserved region and is part of the active site of the enzyme. The cysteine residue at position seven was predominantly labeled and suggested to be the site modified by arylazido-beta-alanyl-NAD+. This cysteine residue corresponds to the Cys-149 in the pig muscle enzyme, which has been shown to be an essential residue for the enzyme activity. The present investigation clearly demonstrates that arylazido-beta-alanyl-NAD+ is a useful photoaffinity probe to characterize the active sites of NAD(H)-dependent enzymes.  相似文献   

11.
Yeast glyceraldehyde-3-phosphate dehydrogenase (GPDH) covalently attached to CNBr-activated Sepharose 4B was shown to be capable of binding soluble yeast phosphoglycerate kinase (PGK) in the course of incubation in the presence of an excess of 1,3-diphosphoglycerate. The association of the matrix-bound and soluble enzymes also occurred if the kinase was added to a reaction mixture in which the immobilized glyceraldehyde-3-phosphate dehydrogenase, NAD, glyceraldehyde-3-phosphate and Pi had been preincubated. Three kinase molecules were bound per a tetramer of the immobilized dehydrogenase and one molecule per a dimer. An immobilized monomer of glyceraldehyde-3-phosphate dehydrogenase was incapable of binding phosphoglycerate kinase. The matrix-bound bienzyme complexes were stable enough to survive extensive washings with a buffer and could be used repeatedly for activity determinations. Experimental evidence is presented to support the conclusion that 1,3-diphosphoglycerate produced by the kinase bound in a complex can dissociate into solution and be utilized by the dehydrogenase free of phosphoglycerate kinase.  相似文献   

12.
Initial rate studies at pH 7.6 with three aldehydes, product inhibition patterns with NADH and dead-end inhibition with adenosine diphosphoribose show that the kinetic mechanism of glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle cannot be ordered, and support an enzyme-substitution mechanism. Deviations from Michaelis-Menten behaviour are consistent with negative interactions in the binding of NAD+ and instability of the species E(NAD)3 and E(NAD)4. Inhibition with large concentrations of phosphate and arsenate indicates competition for a binding site for glyceraldehyde 3-phosphate, and is not found with glyceraldehyde as substrate.  相似文献   

13.
The NH2-terminal amino acid sequence of rat skeletal muscle glyceraldehydephosphate dehydrogenase (D-glyceraldehyde-3-phosphate : NAD+ oxidoreductase(physphorylating), EC 1.2.1.12) was determined to be Val-Lys-Val-Gly-Val-Asn-Gly-Phe-Gly-Arg-Ile-Gly-Arg-Leu-Val-Thr-Arg-Ala-Ala-Phe-Ser-Ser-(-)-(-)--Val-Asx-Ile-Val-Ala-Ile. The presence of Asn instead of Asp in position 6 differentiates this enzyme from other glyceraldehyde-3-phosphate dehydrogenases so far sequenced with the exception of the enzymes isolated from liver. The location of Asn in position 6 has been considered as a specific property of liver glyceraldehyde-3-phosphate dehydrogenase (Kulbe, K.D., Jackson, K.W. and Tang, J. (1975) Biochem. Biophys. Res. Commun. 67, 35--42); this suggestion is not sustained by the results of the present investigation. The amino acid composition of the rat skeletal muscle dehydrogenase demonstrates the unusually low histidine content of this enzyme as compared to other mammalian muscle glyceraldehyde-phosphate dehydrogenases.  相似文献   

14.
Yeast glyceraldehyde-3-phosphate dehydrogenase (glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12) immobilized on CNBr-activated Sepharose 4-B has been subjected to dissociation to obtain matrix-bound dimeric species of the enzyme. Hybridization was then performed using soluble glyceraldehyde-3-phosphate dehydrogenase isolated from rat skeletal muscle. Immobilized hybrid tetramers thus obtained were demonstrated to exhibit two distinct pH-optima of activity characteristic of the yeast and muscle enzymes, respectively. The results indicate that under appropriate conditions the activity of each of the dimers composing the immobilized hybrid tetramer can be studied separately.  相似文献   

15.
Summary Hypotonic human erythrocyte ghosts, devoid of the original glyceraldehyde-3-phosphate dehydrogenase content of the red cell, bind added glyceraldehyde-3-phosphate dehydrogenases, isolated from human erythrocytes, rabbit and pig muscle, as well as rabbit muscle aldolase. There are only slight differences in the affinities towards the various glyceraldehyde-3-phosphate dehydrogenases. On the other hand, glyceraldehyde-3-phosphate dehydrogenases are bound much stronger than aldolase; in an equimolar mixture the former can prevent the binding of the latter, or replace previously bound aldolase at the membrane surface. Binding is always accompanied by the partial inactivation of enzymes, which can be reverted by desorption. Unwashed ghosts rich in hemoglobin seem to have a more pronounced inactivating effect on bound glyceraldehyde-3-phosphate dehydrogenase. In isotonic media ghosts, whether white or unwashed, reseal and do not interact with the enzymes.  相似文献   

16.
Oligonucleotide-directed mutagenesis was employed to produce mutants of the glyceraldehyde-3-phosphate dehydrogenase (GAPDH) of Escherichia coli and Bacillus stearothermophilus. Three different mutants proteins--His176----Asn, Cys149----Ser, Cys149----Gly--were isolated from one or both of the enzymes. The study of the properties of these mutants has shown that Cys149 is clearly responsible for the information of a charge-transfer transition, named the Racker band, observed during the NAD+ binding to apoGAPDH. This result excludes a similarity between the Racker band and the charge-transfer transition observed following the alkylation of GAPDH by 3-chloroacetyl pyridine-adenine dinucleotide.  相似文献   

17.
Glyceraldehyde 3-phosphate dehydrogenase is a tetramer of four chemically identical subunits which requires the cofactor nicotinamide adenine dinucleotide (NAD) for activity. The structure of the holo-enzyme from Bacillus stearothermophilus has recently been refined using X-ray data to 2.4 A resolution. This has facilitated the structure determination of both the apo-enzyme and the enzyme with one molecule of NAD bound to the tetramer. These structures have been refined at 4 A resolution using the constrained-restrained parameter structure factor least-squares refinement program CORELS. When combined with individual atomic temperature factors from the holo-enzyme, these refined models give crystallographic R factors of 30.2% and 30.4%, respectively, for data to 3 A resolution. The apo-enzyme has 222 molecular symmetry, and the subunit structure is related to that of the holo-enzyme by an approximate rigid-body rotation of the coenzyme binding domain by 4.3 degrees with respect to the catalytic domains, which form the core of the tetramer. The effect of this rotation is to shield the coenzyme and active site from solvent in the holo-enzyme. In addition to the rigid-body rotation, there is a rearrangement of several residues involved in NAD binding. The structure of the 1 NAD enzyme is asymmetric. The subunit which contains the bound NAD adopts a conformation very similar to that of a holo-enzyme subunit, while the other three unliganded subunits are very similar to the apo-enzyme conformation. This result provides unambiguous evidence for ligand-induced sequential conformational changes in B. stearothermophilus glyceraldehyde 3-phosphate dehydrogenase.  相似文献   

18.
D Eby  M E Kirtly 《Biochemistry》1976,15(10):2168-2171
Using NAD analogues as ligands, the structural requirements for negative cooperativity in binding to rabbit muscle glyceraldehyde-3-phosphate dehydrogenase were examined. Although the affinity of nicotinamide hypoxanthine dinucleotide is considerably lower than that of NAD+, it also binds to the enzyme with negative cooperatively. Two pairs of nicotinamide hypoxanthine dinucleotide binding sitess were distinguished, one pair having an affinity for the analogue which is 15 times that of the second pair. Negative cooperativity is also found in the Km values for the analogue. Thus modification of the adenine ring of NAD+ to hypoxanthine does not abolish negative cooperativity in coenzyme binding. Adenosine diphosphoribose binding to the same enzyme shows neither positive nor negative cooperativity, indicating that cooperativity apparently requires an intact nicotinamide ring in the coenzyme structure, under the conditions of these experiments. Occupancy of the nicotinamide subsite of the coenzyme binding site is not necessary for half-of-sites reactivity of alkylating or acylating compounds (Levitzki, A. (1974), J. Mol, Biol. 90, 451-458). However, it can be important in the negative cooperativity in ligand binding, as illustrated by adenosine diphosphoribose which fails to exhibit negative cooperativity. Occupancy of the adenine subsite by adenine is important for stabilization of the enzyme against thermal denaturation. Whether the stabilization is due to an altered conformation of the subunits or stabilization of the preexisting structure of the apoenzyme cannot be determined from these studies. However, nicotinamide hypoxanthine dinucleotide does not contribute to enzyme stability although it serves as a substrate and shows negative cooperativity.  相似文献   

19.
The binding of a spin-labeled AMP analog to tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle is described. The spin label, perdeuterated and 15N-substituted 4-amino-2,2,6,6-tetramethylpiperidine-1-oxyl, was attached to C-8 of AMP (C8-SL-AMP). Up to 8 equivalents of C8-SL-AMP bind per enzyme tetramer, i.e., 2 per monomer. Combining sites are the adenine subsite of the coenzyme-binding domain and the phosphate site. Glyceraldehyde 3-phosphate causes a conformational change in the enzyme that brings C8-SL-AMP molecules bound to adjacent R-axis-related subunits closer to one another by 0.2-0.3 nm and allows for spin-spin interaction between the nitroxide radicals. Similar, but less pronounced structural changes take place upon lowering the pH from 8 to 7. Addition of a single equivalent of NAD+ to a complex of the enzyme with 7.6 equivalents of C8-SL-AMP leads to the release of almost 4 C8-SL-AMP molecules. This supports our previous findings that binding of just one NAD+ molecule induces conformational changes in all four subunits.  相似文献   

20.
Abstract— Cat sciatic nerves were exposed to iodoacetate for a period of 5–10 min and after washing out the iodoacetate, the enzymes, glyceraldehyde-3-phosphate dehydrogenase ( d -glyceraldehyde-3-phosphate: NAD oxidoreductase (phosphorylating); EC 1.2.1.12) and lactate dehydrogenase ( l -lactate: NAD oxidoreductase; EC 1.1.1.27) were extracted from the high-speed supernatant fraction of nerve homogenates. Concentrations of iodoacetate as low as 2.5 m m could completely block activity of glyceraldehyde-3-phosphate dehydrogenase but had no effect on lactate dehydrogenase. These findings are in accord with the classical concept shown earlier for muscle that iodoacetate blocks glycolysis by its action on glyceraldehyde-3-phosphate dehydrogenase. A complete block of activity of the enzyme was found after treatment with 2 to 5 m m -iodoacetate for a period of 10 min and such blocks were irreversible for at least 3 h. Glyceraldehyde-3-phosphate dehydrogenase activity was NAD specific, with NADP unable to substitute for NAD. The results are discussed in relation to the effect of iodoacetate in blocking glycolysis and in turn the fast axoplasmic transport of materials in mammalian nerve.  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