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.
Koningic acid, a sesquiterpene antibiotic, is a specific inhibitor of the enzyme glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12). In the presence of 3 mM of NAD+, koningic acid irreversibly inactivated the enzyme in a time-dependent manner. The pseudo-first-order rate constant for inactivation (kapp) was dependent on koningic acid concentration in saturate manner, indicating koningic acid and enzyme formed a reversible complex prior to the formation of an inactive, irreversible complex; the inactivation rate (k 3) was 5.5.10(-2) s-1, with a dissociation constant for inactivation (Kinact) of 1.6 microM. The inhibition was competitive against glyceraldehyde 3-phosphate with a Ki of 1.1 microM, where the Km for glyceraldehyde 3-phosphate was 90 microM. Koningic acid inhibition was uncompetitive with respect to NAD+. The presence of NAD+ accelerated the inactivation. In its absence, the charcoal-treated NAD+-free enzyme showed a 220-fold decrease in apparent rate constant for inactivation, indicating that koningic acid sequentially binds to the enzyme next to NAD+. The enzyme, a tetramer, was inactivated when maximum two sulfhydryl groups, possibly cysteine residues at the active sites of the enzyme, were modified by the binding of koningic acid. These observations demonstrate that koningic acid is an active-site-directed inhibitor which reacts predominantly with the NAD+-enzyme complex.  相似文献   

2.
Transient kinetic methods have been used to study the influence of NAD(+) on the rate of elementary processes of the reversible oxidative phosphorylation of d-glyceraldehyde 3-phosphate catalysed by d-glyceraldehyde 3-phosphate dehydrogenase. In the pH range 5-8 NAD(+) is bound to the enzyme during the following elementary processes of the mechanism: phosphorolysis of the acyl-enzyme, its formation from 1,3-diphosphoglycerate and the enzyme and the formation and breakdown of the glyceraldehyde 3-phosphate-enzyme complex. The rates of these four elementary processes only equal or exceed the turnover rate of the enzyme when NAD(+) is bound and are as much as 10(4) times the rates in the absence of NAD(+). Autocatalysis of the reductive dephosphorylation of 1,3-diphosphoglycerate occurs when glyceraldehyde 3-phosphate release is rate determining because NAD(+) is a reaction product. An important feature of the enzyme mechanism is that the negative-free-energy change of a chemical reaction, acyl-enzyme formation, is linked in a simple way to the positive-free-energy change of a dissociation reaction, NAD(+) release.  相似文献   

3.
The formation of binary complexes between sturgeon apoglyceralddhyde-3-phosphate dehydrogenase, coenzymes (NAD+ and NADH) and substrates (phosphate, glyceraldehyde 3-phosphate and 1,3-bisphosphoglycerate) has been studied spectrophotometrically and spectrofluorometrica-ly. Coenzyme binding to the apoenzyme can be characterized by several distinct spectroscopic properties: (a) the low intensity absorption band centered at 360 nm which is specific of NAD+ binding (Racker band); (b) the quenching of the enzyme fluorescence upon coenzyme binding; (c) the quenching of the fluorescence of the dihydronicotinamide moiety of the reduced coenzyme (NADH); (D) the hypochromicity and the red shift of the absorption band of NADH centered at 338 nm; (e) the coenzyme-induced difference spectra in the enzyme absorbance region. The analysis of these spectroscopic properties shows that up to four molecules of coenzyme are bound per molecule of enzyme tetramer. In every case, each successively bound coenzyme molecule contributes identically to the total observed change. Two classes of binding sites are apparent at lower temperatures for NAD+ Binding. Similarly, the binding of NADH seems to involve two distinct classes of binding sites. The excitation fluorescence spectra of NADH in the binary complex shows a component centered at 260 nm as in aqueous solution. This is consistent with a "folded" conformation of the reduced coenzyme in the binary complex, contradictory to crystallographic results. Possible reasons for this discrepancy are discussed. Binding of phosphorylated substrates and orthophosphate induce similar difference spectra in the enzyme absorbance region. No anticooperativity is detectable in the binding of glyceraldehyde 3-phosphate. These results are discussed in light of recent crystallographic studies on glyceraldehyde-3-phosphate dehydrogenases.  相似文献   

4.
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   

5.
Bovine lens aldehyde dehydrogenase. Kinetics and mechanism.   总被引:3,自引:2,他引:1       下载免费PDF全文
Bovine lens cytoplasmic aldehyde dehydrogenase exhibits Michaelis-Menten kinetics with acetaldehyde, glyceraldehyde 3-phosphate, p-nitrobenzaldehyde, propionaldehyde, glycolaldehyde, glyceraldehyde, phenylacetylaldehyde and succinic semialdehyde as substrates. The enzyme was also active with malondialdehyde, and exhibited an esterase activity. Steady-state kinetic analyses show that the enzyme exhibits a compulsory-ordered ternary-complex mechanism with NAD+ binding before acetaldehyde. The enzyme was inhibited by disulfiram and by p-chloromercuribenzoate, and studies with with mercaptans indicated the involvement of thiol groups in catalysis.  相似文献   

6.
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   

7.
Summary Partially purified flounder muscle (Pseudopleuronectus americanus) glyceraldehyde 3-phosphate dehydrogenase was immobilized on cyanogen bromide-activated Sepharose. The catalytic properties of the immobilized preparation were studied to determine if immobilization alters the kinetic properties of the native holoenzyme. The results indicate that the pH activity profile of immobilized glyceraldehyde 3-phosphate dehydrogenase did not differ from that of the native enzyme. The Michaelis constants (Km) for NAD and glyceraldehyde 3-phosphate were somewhat altered. The enzyme stability toward various inactivation treatments in the presence and absence of NAD was characterized and compared to that of he native enzyme. When either form of the enzyme was incubated with urea at concentrations greater than 2m, inactivation occurred very rapidly. Incubation in 0.1% trypsin for 60 minutes decreased the activity of immobilized glyceraldehyde 3-phosphate dehydrogenase by 45% and of the native soluble enzyme by 70%. The immobilized enzyme also exhibited considerably more stability than the native soluble enzyme when exposed to a temperature of 50° or to 20 mm ATP. In all cases NAD either greatly reduced the rate of inactivation or completely protected the enzyme from inactivation.  相似文献   

8.
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   

9.
The route of ethanol formation in Zymomonas mobilis   总被引:5,自引:2,他引:5  
1. Enzymic evidence supporting the operation of the Entner-Doudoroff pathway in the anaerobic conversion of glucose into ethanol and carbon dioxide by Zymomonas mobilis is presented. 2. Cell extracts catalysed the formation of equimolar amounts of pyruvate and glyceraldehyde 3-phosphate from 6-phosphogluconate. Evidence that 3-deoxy-2-oxo-6-phosphogluconate is an intermediate in this conversion was obtained. 3. Cell extracts of the organism contained the following enzymes: glucose 6-phosphate dehydrogenase (active with NAD and NADP), ethanol dehydrogenase (active with NAD), glyceraldehyde 3-phosphate dehydrogenase (active with NAD), hexokinase, gluconokinase, glucose dehydrogenase and pyruvate decarboxylase. Extracts also catalysed the overall conversion of glycerate 3-phosphate into pyruvate in the presence of ADP. 4. Gluconate dehydrogenase, fructose 1,6-diphosphate aldolase and NAD-NADP transhydrogenase were not detected. 5. It is suggested that NAD is the physiological electron carrier in the balanced oxidation-reduction involved in ethanol formation.  相似文献   

10.
Tetrameric rabbit muscle glyceraldehyde 3-phosphate dehydrogenase (GAPDH; EC 1.2.1.12) binds successively four molecules of its cofactor (NAD+) with affinities of ca 10(11) M(-1), 10(9) M(-1), 10(7) M(-1), and 10(5) M(-1). The reduction in the dynamics of the protein is greatest upon binding the first NAD+ molecule. Smaller reductions then occur upon binding the second and third NAD+ molecules, and the fourth NAD+ molecule binds without dynamic change. Reduction of the GAPDH dynamics, with consequent improvements in its internal bonding, can account for the increase in NAD+ binding affinity from 10(5) M(-1) to 10(11) M(-1). Evidence is provided that comparable fractions of the binding energy of other ligands, and of the catalytic efficiency of enzymes, may be derived in the same way.  相似文献   

11.
The glyceraldehyde 3-phosphate dehydrogenase activity of extracts from heterotrophic Scenedesmus obliquus was linked predominantly to NADH. However, on DEAE-cellulose chromatography the enzyme was eluted by a gradient of phosphate in a form characterized by high NADPH-dependent glyceraldehyde 3-phosphate dehydrogenase activity. This interconversion of enzyme forms could be prevented by the presence of NAD during DEAE-cellulose chromatography.High concentrations of phosphate stimulated the NADPH-dependent activity of the purified enzyme at the expense of activity linked to NADH and these changes were associated with depolymerization of a hexadecamer to a tetramer. The effect of phosphate on the rates of increase in NADPH-dependent activity and of a decrease in activity linked to NADH was cooperative with a Hill coefficient of 3.2. The inversely related changes in coenzyme specificity were inhibited to the same extent by NAD and the response to this ligand was anticooperative. These findings imply a strictly inverse proportional relationship between the rates of change of NADH and NADPH-linked activity. In the presence of dithiothreitol, low concentrations of phosphate promoted NADPH-dependent activity by stabilising the unstable tetrameric form produced from the hexadecamer by the thiol.These phenomena are discussed in relation to a general mechanism for the in vivo promotion of NADPH-dependent glyceraldehyde 3-phosphate dehydrogenase activity.  相似文献   

12.
Glyceraldehyde has been known to be an insulin secretagogue for more than 15 years. It has been (reasonably) assumed that glyceraldehyde enters the glycolytic pathway via its phosphorylation by ATP to form glyceraldehyde phosphate, a reaction catalyzed by the enzyme triokinase, and that subsequent metabolism is identical to that of glucose. glucose. However, up to now there have been no studies verifying the presence of triokinase in the pancreatic beta cell. We report here that (1) the activity of triokinase in pancreatic islets is very low, indicating that the activity is intrinsically low and/or the enzyme was rapidly inactivated during the preparation of tissue for assay; (2) the activity is much lower than glucose phosphorylating activity (hexokinase plus glucokinase) in islets, even though glyceraldehyde is a more efficient insulin secretagogue than glucose; (3) glyceraldehyde phosphate dehydrogenase from pancreatic islets can use glyceraldehyde as a substrate in place of glyceraldehyde phosphate (the Vmax of glyceraldehyde phosphate dehydrogenase from islets when glyceraldehyde is the substrate is 20-fold that of triokinase when glyceraldehyde is the substrate); and (4) the Km of glyceraldehyde phosphate dehydrogenase with respect to glyceraldehyde (4.8 mM) is similar to the concentration of glyceraldehyde that gives one-half maximal rates of insulin release from pancreatic islets, whereas the Km of triokinase with respect to glyceraldehyde is much lower (less than 50 microM). These data suggest that besides stimulating insulin release in islets via its entering metabolism by phosphorylation to glyceraldehyde phosphate in the triokinase reaction, glyceraldehyde could be phosphorylated by Pi in the glyceraldehyde phosphate dehydrogenase reaction to form glycerate 1-phosphate which is probably unmetabolizable in islets. The second reaction could drastically increase the NADH/NAD ratio in islets without providing substrates for hydrogen shuttles that reoxidize cytosolic NADH. Since an increased NAD(P)H/NAD(P) ratio is believed to be a key part of the signal for insulin release, such a mechanism would explain the potent insulinotropism of glyceraldehyde in short-term experiments. In addition, the formation of unmetabolizable acids may explain the toxic effects of long-term exposure of islets to glyceraldehyde and why glyceraldehyde causes the beta cell to become acidic, whereas glucose does not.  相似文献   

13.
The regulation of CO(2) assimilation by intact spinach (Spinacia oleracea) chloroplasts by exogenous NADP-linked nonreversible d-glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.9) was investigated. This dehydrogenase mediated a glyceraldehyde 3-phosphate/glycerate 3-phosphate shuttle for the indirect transfer of NADPH from chloroplast to the external medium. The rate of NADPH formation in the medium reflected glyceraldehyde 3-phosphate efflux from the chloroplast. Increasing enzyme concentrations stimulated NADP reduction and, in turn, CO(2) fixation. Pyrophosphate increased CO(2) fixation by apparently inhibiting glyceraldehyde 3-phosphate efflux. Increasing the glycerate 3-phosphate concentration above 0.1 mm stimulated glyceraldehyde 3-phosphate efflux but inhibited CO(2) fixation. Addition of up to 0.5 mm orthophosphate enhanced both glyceraldehyde 3-phosphate efflux and CO(2) fixation while each was inhibited by higher orthophosphate concentrations. The mechanism by which the extent of glyceraldehyde 3-phosphate efflux regulated the rate of CO(2) fixation in chloroplasts was discussed.  相似文献   

14.
Tetrameric D-glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12) isolated from rabbit skeletal muscle was covalently bound to CNBr-activated Sepharose 4B via a single subunit. Catalytically active immobilized dimer and monomeric forms of the enzyme were prepared after urea-induced dissociation of the tetramer. A study of the coenzyme-binding properties of matrix-bound tetrameric, dimeric and monomeric species has shown that: (1) an immobilized tetramer binds NAD+ with negative cooperativity, the dissociation constants being 0.085 microM for the first two coenzyme molecules and 1.3 microM for the third and the fourth one; (2) coenzyme binding to the dimeric enzyme form also displays negative cooperativity with Kd values of 0.032 microM and 1.1 microM for the first and second sites, respectively; (3) the binding of NAD+ to a monomer can occur with a dissociation constant of 1.6 microM which is close to the Kd value for low-affinity coenzyme binding sites of the tetrameric or dimeric enzyme forms. In the presence of NAD+ an immobilized monomer acquires a stability which is not inferior to that of a holotetramer. The catalytic properties of monomeric and tetrameric enzyme forms were compared and found to be different under certain conditions. Thus, the monomers of rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase displayed a hyperbolic kinetic saturation curve for NAD+, whereas the tetramers exhibited an intermediary plateau region corresponding to half-saturating concentrations of NAD+. At coenzyme concentrations below half-saturating a monomer is more active than a tetramer. This difference disappears at saturating concentrations of NAD+. Immobilized monomeric and tetrameric forms of D-glyceraldehyde-3-phosphate dehydrogenase from baker's yeast were also used to investigate subunit interactions in catalysis. The rate constant of inactivation due to modification of essential arginine residues in the holoenzyme decreased in the presence of glyceraldehyde 3-phosphate, probably as a result of conformational changes accompanying catalysis. This effect was similar for monomeric and tetrameric enzyme forms at saturating substrate concentrations, but different for the two enzyme species under conditions in which about one-half of the active centers remained unsaturated. Taken together, the results indicate that association of D-glyceraldehyde-3-phosphate dehydrogenase monomers into a tetramer imposes some constraints on the functioning of the active centers.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

15.
An NAD-dependent glycerol 3-phosphate dehydrogenase (sn-glycerol 3-phosphate: NAD oxidoreductase; EC 1.1.1.8) has been purified from spinach leaves by a three-step procedure involving ion-exchange, gel filtration, and affinity chromatography. The enzyme has been purified over 10,000-fold to a specific activity of 38. It has a molecular weight of approximately 63,500. The pH optimum for the reduction of dihydroxyacetone phosphate is 6.8 and for glycerol 3-phosphate oxidation it is 9.5. During dihydroxyacetone phosphate reduction hyperbolic kinetics were observed when either NADH or dihydroxyacetone phosphate was the variable substrate, but concentrations of NADH greater than 150 μm were inhibitory. Michaelis constants were 0.30–0.35 mm for dihydroxyacetone phosphate and 0.01 mm for NADH. Glycerol 3-phosphate oxidation obeyed Michaelis-Menten kinetics with a Km of 0.19 mm for NAD and 1.6 mm for glycerol 3-phosphate. The enzyme was specific for those substrates, and dihydroxyacetone, glyceraldehyde, glyceraldehyde 3-phosphate, NADPH, NADP, and glycerol were not utilized. The spinach leaf enzyme appears to be in the cytoplasm and probably functions for the production of glycerol 3-phosphate from dihydroxyacetone phosphate.  相似文献   

16.
The kinetic mechanisms of the NAD- and NADP-linked reactions catalyzed by glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides were examined using product inhibition, dead-end inhibition and alternate substrate experiments. The results are consistent with a steady-state random mechanism for the NAD-linked and an ordered, sequential mechanism with NADP+ binding first for the NADP-linked reaction. Thus, the enzyme can bind NADP+, NAD+, and glucose 6-phosphate, but the enzyme-glucose 6-phosphate complex can react only with NAD+, not with NADP+. This affects the rate equation for the NADP-linked reaction by introducing a term for a dead-end enzyme-glucose 6-phosphate complex. The kinetic mechanisms represent revisions of those proposed previously (C. Olive, M.E. Geroch, and H.R. Levy, 1971, J. Biol. Chem. 246, 2047-2057) and provide a kinetic basis for the regulation of coenzyme utilization of the enzyme by glucose 6-phosphate concentration (H.R. Levy, and G.H. Daouk, 1979, J. Biol. Chem. 254, 4843-4847) and NADPH/NADP+ concentration ratios (H.R. Levy, G.H. Daouk, and M.A. Katopes, 1979, Arch, Biochem. Biophys. 198, 406-413). The kinetic mechanisms were found to be the same at pH 6.2 and pH 7.8. The kinetics of ATP inhibition of the NAD- and NADP-linked reactions were examined at pH 6.2 and pH 7.8. The results are interpreted in terms of ATP addition to binary enzyme-coenzyme and enzyme-glucose 6-phosphate complexes.  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.
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   

19.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides is inactivated by trypsin, chymotrypsin, pronase E, thermolysin, 4.0 M urea, and by heating to 49 degrees C. It is protected, to varying degrees, against all these forms of inactivation by glucose 6-phosphate, NAD+, and NADP+. When these ligands are present at 10 times their respective KD concentrations, protection by NAD+ or glucose 6-phosphate is substantially greater than protection by NADP+. A detailed analysis was undertaken of the protective effects of these ligands, at varying concentrations, on proteolysis of glucose-6-phosphate dehydrogenase by thermolysin. This study confirmed the above conclusion and permitted calculation of KD values for NAD+, NADP+, and glucose 6-phosphate that agree with such values determined by independent means. For NADP+, two KD values, 6.1 microM and 8.0 mM, can be derived, associated with protection against thermolysin by low and high NADP+ concentrations, respectively. The former value is in agreement with other determinations of KD and the latter value appears to represent binding of NADP+ to a second site which causes inhibition of catalysis. A Ki value of 10.5 mM for NADP+ was derived from inhibition studies. The principal conclusion from these studies is that NAD+ binding to L. mesenteroides glucose-6-phosphate dehydrogenase results in a larger global conformational change of the enzyme than does NADP+ binding. Presumably, a substantially larger proportion of the free energy of binding of NAD+, compared to NADP+, is used to alter the enzyme's conformation, as reflected in a much higher KD value. This may play an important role in enabling this dual nucleotide-specific dehydrogenase to accommodate either NAD+ or NADP+ at the same binding site.  相似文献   

20.
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   

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