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 spatial arrangement of coenzyme NAD+ in remote and adjacent binding sites in various stoichiometric complexes with tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle was examined via EPR spectroscopy. An adenosine N6-15N,2H17 spin-labeled derivative of coenzyme NAD+ (SL-NAD+) was chemically synthesized for this work. The spectral simplifications and narrow line widths afforded by 15N and 2H substitution enabled experimental EPR spectra to be deconvoluted into their three component spectra: (a) unbound coenzyme, (b) bound coenzyme without adjacent site occupied, and (c) bound coenzyme with adjacent site occupied. Binding of SL-NAD+ in adjacent active centers of R axis-related subunits resulted in resolved dipolar interactions which characterized intersubunit distances. Binding to distant subunits related by the P and Q axes gave no dipolar interaction. Once the first NAD+ site was occupied, EPR spectra at various stoichiometries provided evidence for nonpreferential spatial binding of SL-NAD+ to the three unoccupied sites. EPR spectral simulations indicated a separation of 12.8 A for the unpaired electrons of spin label moieties of R axis-related coenzymes. Molecular modeling based on x-ray crystallographic data predicted 11-13 A. The angles and distance relating to interacting spin-labels were calculated from atomic coordinates based on molecular modeling of both anti-anti and anti-syn (adenine-ribose) conformations of SL-NAD+. Computer-generated line shapes indicated best agreement with experimental EPR results when the anti-anti geometry was employed. Comparison of EPR spectra from soluble and ammonium sulfate-precipitated enzymes indicated that the NAD+-binding domains are positioned equivalently in the two physical states. Since the observed dipolar line shapes are critically dependent on the distance and geometry relating to the interacting SL-NAD+, these data provide direct evidence for a high degree of conservation of quaternary structure of the enzyme in the hydrated crystalline state. Studies on the enzyme isolated from human erythrocytes also indicated a close correlation with the rabbit muscle enzyme in both the arrangement of NAD+-binding domains and negative cooperativity of coenzyme binding.  相似文献   

2.
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   

3.
Glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12) was isolated from a sturgeon, Huso huso, from the Caspian Sea. It is closely related to the enzyme from a Pacific sturgeon, Acipenser transmontanus, with respect to amino acid composition, steady-state kinetics and coenzyme binding. The latter, as studied by means of a spin-labeled derivative of NAD+, is negatively cooperative exhibiting a Hill coefficient of 0.84 at 12 degrees C. Two derivatives of NAD+ spin-labeled at N6 or C8 of the adenine ring were found to be active coenzymes with maximum velocities reaching 35 or 45% of the value for NAD+ itself. When more than two equivalents of either spin-labeled NAD+ are bound to the enzyme spin-spin interactions are observed in the ESR spectra. Distances between the nitroxide radicals (8--9 A) calculated from the observed splittings are in excellent agreement with data predicted from the crystal structure of the lobster enzyme when the coenzyme is bound in an anti-conformation of the adenine moiety about the glycosidic bond to all four subunits.  相似文献   

4.
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   

5.
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   

6.
Glyceraldehyde 3-phosphate dehydrogenase (D-glyceraldehyde-3-phoshate:nicotinamide adenine dinucleotide oxidoreductase (phosphorylating), EC 1.2.1.12) forms a complex with 3-pyridinealdehyde-NAD which survives precipitation with 7% perchloric acid. The molar ratio bound 3-pyridinealdehyde-NAD to the enzyme is 2.5 to 2.9. Lactate, malate, and alcohol dehydrogenases do not form acid-precipitable complexes with 3-pyridinealdehyde-NAD. 3-Pyridinealdehyde-deamino-NAD or glyceraldehyde 3-phosphate also forms an acid-stable complex with glyceraldehyde 3-phosphate dehydrogenase; however, NAD, 3-acetylpyridine-NAD, or thionicotinamide-NAD does not produce an acid-stable complex. Incubation of the glyceraldehyde 3-phosphate dehydrogenase with glyceraldehyde 3-phosphate, acetyl phosphate, iodoacetic acid, or iodosobenzoate inhibits the formation of the acid-stable complex with 3-pyridinealdehyde-NAD. Glyceraldehyde 3-phosphate or 3-pyridinealdehyde-NAD also prevents carboxymethylation of the active site cysteine-149 by[14-C]iodoacetic acid. These studies indicate that the aldehyde group of 3-pyridinealdehyde-NAD forms a thiohemiacetal linkage with cysteine-149 which is the substrate binding site for the dehydrogenase reaction. These findings may account for the fact that 3-pyridinealdehyde-NAD strongly inhibits the dehydrogenase and esterase activities of 3-pyridinealdehyde-NAD forms a thiohemiacetal linkage with cysteine-149 which is the substrate binding site for the dehydrogenase reaction. These findings may account for the fact that 3-pyridinealdehyde-NAD strongly inhibits the dehydrogenase and esterase activities of glyceraldehyde 3-phosphate dehydrogenase which require reduced cysteine-149. However, the analogue does not inhibit the acetyl phosphates activity of the enzyme for which the active site sulfhydryl residues must be oxidized.  相似文献   

7.
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   

8.
A series of NAD+ analogues, modified on the pyridinium ring, have been tested for their enzymic properties in reactions with D-glyceraldehyde-3-phosphate dehydrogenase form sturgeon muscle, rabbit muscle and Bacillus stearothermophilus. The observed activity, inhibition and binding data are correlated to the structure of the enzyme and coenzyme analogue by model building on a Vector General interactive graphic display system using coordinates from the B. stearothermophilus holoenzyme structure. Most of the analogues with substituents in the pyridinium-3 position could be bound to glyceraldehyde-3-phosphate dehydrogenase, either in manner similar to NAD+ or in a completely different way with the substituted pyridinium ring rotated 110 degrees or more around the glycosidic bond. This indicates different possible modes of binding of NAD+ analogues within the pyridinium binding subsite. Analogues with substituents in the pyridinium-4 position are shown to be weakly bound to glyceraldehyde-3-phosphate dehydrogenase. This is explained by a strong interaction of the substituent in the 4 position with the residues Asn-313 and Cys-149.  相似文献   

9.
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   

10.
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   

11.
平衡柱层析法测得每分子龙虾肌羧甲基化甘油醛-3磷酸脱氢酶能结合3.9分子NAD~+,而每分子光照酶则只能结合2分子NAD~+。 由蛋白荧光淬灭法得到,在25℃、pH7.0的磷酸盐缓冲液中,全酶、羧甲基酶及光照酶与NAD~+结合时均呈负协同性。  相似文献   

12.
A new method of calculation, based on a direct fitting of the protein fluorescence intensity observed upon coenzyme binding (H.-P. Lutz, unpublished results), is used to study the negative cooperative behavior of glyceraldehyde-3-phosphate dehydrogenase from sturgeon muscle. The calculation procedure simultaneously elaborates data obtained for four different protein concentrations, and it is able to compare different models by computing the minimal and critical sum of squares. Using this approach, it is shown that the induced-fit model [Koshland, D. E., Jr., Nemethy, G., & Filmer, D. (1966) Biochemistry 5,365] and the dimer of dimer model [Malhotra, O. P., & Bernhard, S. A. (1968) J. Biol. Chem. 243, 1243-1252] can both be applied for explaining the negative cooperativity observed upon coenzyme binding to sturgeon glyceraldehyde-3-phosphate dehydrogenase. In addition to the progressive modification of the binding affinity during ligand binding, different maximal fluorescence quenchings for the binding steps must be postulated; and furthermore, the binding capability decreases by decreasing the protein concentration. The fact that the induced-fit model can also be applied is rather in contradiction with the view generally accepted of a dimer of dimer structure of sturgeon glyceraldehyde-3-phosphate dehydrogenase. By use of the same approach, nicotinamide 8-bromoadenine dinucleotide is shown to bind to glyceraldehyde-3-phosphate dehydrogenase from sturgeon in a negative cooperative manner.  相似文献   

13.
Experimental conditions favouring the dissociation of tetrameric rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase into active monomers were elaborated. The urea-induced dissociation of the tetramer was shown to be a stepwise process (in 2 M urea only dimers are formed; an increase in urea concentration up to 3 M causes the splitting of the dimers into monomers). The specific activity of immobilized monomers in the glyceraldehyde-3-phosphate oxidation reaction does not differ from that of the parent immobilized tetrameric form. The tetrameric enzyme molecule binds the coenzyme with a negative cooperativity (the first two NAD+ molecules bind with KD below 0.1 microM; for the third and fourth molecules the dissociation constant was determined to be equal to 5.5 +/- 1.5 microM (50 mM medinal buffer, 10 mM sodium phosphate, pH 8.2). The cooperativity of NAD+ binding is preserved in the immobilized preparation of tetrameric dehydrogenase. The immobilized monomers bind NAD+ with KD of 1.6 +/- 1.0 microM. The experimental results are consistent with the hypothesis according to which the association of catalytically active subunits into a tetramer changes their coenzyme-binding properties in such a way that the first two NAD+ molecules bind more firmly to a tetramer than to a monomer, whereas the third and the fourth NAD+ molecules bind less firmly.  相似文献   

14.
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   

15.
Ultraviolet resonance Raman (UVRR) spectra, with 260-nm excitation, are reported for oxidized and reduced nicotinamide adenine dinucleotides (NAD+ and NADH, respectively). Corresponding spectra are reported for these coenzymes when bound to the enzymes glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and liver and yeast alcohol dehydrogenases (LADH and YADH). The observed differences between the coenzyme spectra are interpreted in terms of conformation, hydrogen bonding, and general environment polarity differences between bound and free coenzymes and between coenzymes bound to different enzymes. The possibility of adenine protonation is discussed. UVRR spectra with 220-nm excitation also are reported for holo- and apo-GAPDH (GAPDH-NAD+ and GAPDH alone, respectively). In contrast with the 260-nm spectra, these show only bands due to vibrations of aromatic amino acid residues of the protein. The binding of coenzyme to GAPDH has no significant effect on the aromatic amino acid bands observed. This result is discussed in the light of the known structural change of GAPDH on binding coenzyme. Finally, UVRR spectra with 240-nm excitation are reported for GAPDH and an enzyme-substrate intermediate of GAPDH. Perturbations are reported for tyrosine and tryptophan bands on forming the acyl enzyme.  相似文献   

16.
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   

17.
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   

18.
The fluorescence decay mechanism of 1, N6-ethenoadenosine diphosphoribose bound to rabbit muscle glyceraldehyde 3-phosphate dehydrogenase markedly differs from that of the intact coenzyme analog (εNAD+) bound to the same enzyme. In the latter case the fluorescence is partially quenched by interactions between the ethenoadenine ring and amino acid residues in its binding site. Binding of the nicotinamide moiety of the coenzyme thus affects the relative orientation of the adenine ring within its binding site leading to the quenching interactions. The interactions of the adenine group with its binding site induce conformational changes in the enzyme which affect the binding of additional coenzyme molecules. The nicotinamide base thus determines, indirectly, the negative cooperativity found in NAD+ binding.  相似文献   

19.
D-beta-Hydroxybutyrate dehydrogenase (D-3-hydroxybutyrate:NAD+ oxidoreductase, EC 1.1.1.30) is a lipid-requiring enzyme which specifically requires phosphosphatidylcholine for enzymic activity. The phosphatidylcholine modifies the binding and orientation of the coenzyme, NAD(H), with respect to the enzyme. In the present study, two derivatives of NAD, spin-labeled either at N-6 or C-8 of the adenine ring, were found to be active as coenzyme. The binding affinity of NADH to the enzyme was opitimized by increasing the salt concentration and increasing the pH from 6 to 8, with the pK at 6.8. Monomethylmalonate, a substrate analogue, was found to enhance NADH binding (Kd is reduced from 4 to 1 microM). Sulfite strongly enhances the binding of NAD+ via the enzyme-catalyzed formation of an adduct of sulfite with the nucleotide; the Kd for binding of NAD-sulfite is in the micromolar range, whereas NAD+ binding is more than a magnitude weaker. The binding of spin-labeled NAD(H) was further characterized by EPR spectroscopy. Increased sensitivity and resolution were obtained with the use of NAD(H) analogues perdeuterated in the spin-label moiety. For these analogues bound to D-beta-hydroxybutyrate dehydrogenase in phospholipid vesicles, EPR studies showed the spin-label moiety to be constrained and revealed two distinct components. Increasing the viscosity of the medium by addition of glycerol affected the EPR spectral characteristics of only the component with the smaller resolved averaged hyperfine splitting. The stage is now set to study motional characteristics of the enzyme, using these spin-labeled probes which mimic the coenzyme.  相似文献   

20.
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   

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