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 hyperthermophilic archaeum Thermoproteus tenax possesses two glyceraldehyde-3-phosphate dehydrogenases differing in cosubstrate specificity and phosphate dependence of the catalyzed reaction. NAD+-dependent glyceraldehyde-3-phosphate dehydrogenase catalyzes the phosphate-independent irreversible oxidation of D-glyceraldehyde 3-phosphate to 3-phosphoglycerate. The coding gene was cloned, sequenced, and expressed in Escherichia coli. Sequence comparisons showed no similarity to phosphorylating glyceraldehyde-3-phosphate dehydrogenases but revealed a relationship to aldehyde dehydrogenases, with the highest similarity to the subgroup of nonphosphorylating glyceraldehyde-3-phosphate dehydrogenases. The activity of the enzyme is affected by a series of metabolites. All effectors tested influence the affinity of the enzyme for its cosubstrate NAD+. Whereas NADP(H), NADH, and ATP reduce the affinity for the cosubstrate, AMP, ADP, glucose 1-phosphate, and fructose 6-phosphate increase the affinity for NAD+. Additionally, most of the effectors investigated induce cooperativity of NAD+ binding. The irreversible catabolic oxidation of glyceraldehyde 3-phosphate, the control of the enzyme by energy charge of the cell, and the regulation by intermediates of glycolysis and glucan degradation identify the NAD+-dependent glyceraldehyde-3-phosphate dehydrogenase as an integral constituent of glycolysis in T. tenax. Its regulatory properties substitute for those lacking in the reversible nonregulated pyrophosphate-dependent phosphofructokinase in this variant of the Embden-Meyerhof-Parnas pathway.  相似文献   

2.
During batch growth of Lactococcus lactis subsp. lactis NCDO 2118 on various sugars, the shift from homolactic to mixed-acid metabolism was directly dependent on the sugar consumption rate. This orientation of pyruvate metabolism was related to the flux-controlling activity of glyceraldehyde-3-phosphate dehydrogenase under conditions of high glycolytic flux on glucose due to the NADH/NAD+ ratio. The flux limitation at the level of glyceraldehyde-3-phosphate dehydrogenase led to an increase in the pool concentrations of both glyceraldehyde-3-phosphate and dihydroxyacetone-phosphate and inhibition of pyruvate formate lyase activity. Under such conditions, metabolism was homolactic. Lactose and to a lesser extent galactose supported less rapid growth, with a diminished flux through glycolysis, and a lower NADH/NAD+ ratio. Under such conditions, the major pathway bottleneck was most probably at the level of sugar transport rather than glyceraldehyde-3-phosphate dehydrogenase. Consequently, the pool concentrations of phosphorylated glycolytic intermediates upstream of glyceraldehyde-3-phosphate dehydrogenase decreased. However, the intracellular concentration of fructose-1,6-bisphosphate remained sufficiently high to ensure full activation of lactate dehydrogenase and had no in vivo role in controlling pyruvate metabolism, contrary to the generally accepted opinion. Regulation of pyruvate formate lyase activity by triose phosphates was relaxed, and mixed-acid fermentation occurred (no significant production of lactate on lactose) due mostly to the strong inhibition of lactate dehydrogenase by the in vivo NADH/NAD+ ratio.  相似文献   

3.
Incubation of glyceraldehyde-3-phosphate dehydrogenase with vinyl sulfones resulted in a pseudo first-order loss of enzyme activity. The selective inactivation of the enzyme by vinyl sulfones is suggested from the structural requirement analysis and the enzyme susceptibility test. The enzyme inactivation was strongly reduced in the presence of NAD or glyceraldehyde-3-phosphate, and the prior treatment of the enzyme with 5,5'-dithio-bis-(2-nitrobenzoic acid) prevented the enzyme from the inactivation by vinyl sulfones (> or = 90%). Moreover, the early rapid phase of inactivation was much more responsive to L-cysteine reactivation, compared with the slower phase. Based on these results, it is proposed that vinyl sulfones inactivate the enzyme by inducing the oxidation of cysteine residue and/or covalent binding to cysteine residue in active site.  相似文献   

4.
(1) The effects of long term treatment with 3-acetylpyridine on the stability of enzymes towards heat and trypsin treatment were studied. (2) In the liver NAD or NADP provided a similar degree of protection against heat inactivation at 55 degrees C for 6-phosphogluconate dehydrogenase (24%), glyceraldehyde-3-phosphate dehydrogenase (24%) and malic enzyme (20%), low level of protection of lactate dehydrogenase (13%) but didn't affect acetylcholinesterase at all. In the muscle, however, there was substantial protection against heat inactivation by coenzyme of glyceraldehyde-3-phosphate dehydrogenase (52%), an intermediate level of protection of lactate dehydrogenase (25%), low level of protection of 6-phosphogluconate dehydrogenase (17%) and malic enzyme (17%) and almost no protection of acetylcholinesterase. (3) In the susceptibility towards trypsin a low but similar degree of protection for dehydrogenases by coenzymes was observed in the liver whereas in the muscle there was substantial protection against trypsin inactivation by NAD of glyceraldehyde-3-phosphate dehydrogenase, an intermediate level of protection of 6-phosphogluconate dehydrogenase and malic enzyme and very little protection of lactate dehydrogenase but no protection of acetylcholinesterase. Among enzymes tested, glyceraldehyde-3-phosphate dehydrogenase showed the greatest protection against heat and trypsin inactivation by NAD. (4) The results suggest that the effect of 3-acetylpyridine treatment on the stability of muscle glyceraldehyde-3-phosphate dehydrogenase appears to be quite specific and selective.  相似文献   

5.
We have previously identified glyceraldehyde-3-phosphate dehydrogenase as an immunoglobulin production stimulating factor (IPSF) which facilitated immunoglobulin production by hybridomas and lymphocytes. The IPSF activity of this enzyme was suppressed by the coexistence of some sorts of nucleotides. We now report that the IPSF effect of GAPDH was suppressed by the coexistence of DNA, the inhibiting effect of degraded DNA being inferior to that of long-chain DNA. Both single-stranded and double-stranded synthetic polyribonucleotides also inhibited the IPSF activity of GAPDH. Moreover, nicotinamide adenine dinucleotide (NAD+) repressed the IPSF effect.  相似文献   

6.
The NADH absorbance spectrum of nicotinoprotein (NADH-containing) alcohol dehydrogenase from Amycolatopsis methanolica has a maximum at 326 nm. Reduced enzyme-bound pyridine dinucleotide could be reversibly oxidized by acetaldehyde. The fluorescence excitation spectrum for NADH bound to the enzyme has a maximum at 325 nm. Upon excitation at 290 nm, energy transfer from tryptophan to enzyme-bound NADH was negligible. The fluorescence emission spectrum (excitation at 325 nm) for NADH bound to the enzyme has a maximum at 422 nm. The fluorescence intensity is enhanced by a factor of 3 upon binding of isobutyramide (Kd = 59 microM). Isobutyramide acts as competitive inhibitor (Ki = 46 microM) with respect to the electron acceptor NDMA (N,N-dimethyl-p-nitrosoaniline), which binds to the enzyme containing the reduced cofactor. The nonreactive substrate analogue trifluoroethanol acts as a competitive inhibitor with respect to the substrate ethanol (Ki = 1.6 microM), which binds to the enzyme containing the oxidized cofactor. Far-UV circular dichroism spectra of the enzyme containing NADH and the enzyme containing NAD+ were identical, indicating that no major conformational changes occur upon oxidation or reduction of the cofactor. Near-UV circular dichroism spectra of NADH bound to the enzyme have a minimum at 323 nm (Deltaepsilon = -8.6 M-1 cm-1). The fluorescence anisotropy decay of enzyme-bound NADH showed no rotational freedom of the NADH cofactor. This implies a rigid environment as well as lack of motion of the fluorophore. The average fluorescence lifetime of NADH bound to the enzyme is 0.29 ns at 20 degreesC and could be resolved into at least three components (in the range 0.13-0.96 ns). Upon binding of isobutyramide to the enzyme-containing NADH, the average excited-state lifetime increased to 1.02 ns and could be resolved into two components (0.37 and 1.11 ns). The optical spectra of NADH bound to nicotinoprotein alcohol dehydrogenase have blue-shifted maxima compared to other NADH-dehydrogenase complexes, but comparable to that observed for NADH bound to horse liver alcohol dehydrogenase. The fluorescence lifetime of NADH bound to the nicotinoprotein is very short compared to enzyme-bound NADH complexes, also compared to NADH bound to horse liver alcohol dehydrogenase. The cofactor-protein interaction in the nicotinoprotein alcohol dehydrogenase active site is more rigid and apolar than that in horse liver alcohol dehydrogenase. The optical properties of NADH bound to nicotinoprotein alcohol dehydrogenase differ considerably from NADH (tightly) bound to UDP-galactose epimerase from Escherichia coli. This indicates that although both enzymes have NAD(H) as nonexchangeable cofactor, the NADH binding sites are quite different.  相似文献   

7.
GapB-encoded protein of Escherichia coli and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) share more than 40% amino acid identity. Most of the amino acids involved in the binding of cofactor and substrates to GAPDH are conserved in GapB-encoded protein. This enzyme shows an efficient non-phosphorylating erythrose-4-phosphate dehydrogenase activity (Zhao, G., Pease, A. J., Bharani, N., and Winkler, M. E. (1995) J. Bacteriol. 177, 2804-2812) but a low phosphorylating glyceraldehyde-3-phosphate dehydrogenase activity, whereas GAPDH shows a high efficient phosphorylating glyceraldehyde-3-phosphate dehydrogenase activity and a low phosphorylating erythrose-4-phosphate dehydrogenase activity. To identify the structural factors responsible for these differences, comparative kinetic and binding studies have been carried out on both GapB-encoded protein of Escherichia coli and GAPDH of Bacillus stearothermophilus. The KD constant of GapB-encoded protein for NAD is 800-fold higher than that of GAPDH. The chemical mechanism of erythrose 4-phosphate oxidation by GapB-encoded protein is shown to proceed through a two-step mechanism involving covalent intermediates with Cys-149, with rates associated to the acylation and deacylation processes of 280 s-1 and 20 s-1, respectively. No isotopic solvent effect is observed suggesting that the rate-limiting step is not hydrolysis. The rate of oxidation of glyceraldehyde 3-phosphate is 0.12 s-1 and is hydride transfer limiting, at least 2000-fold less efficient compared with that of erythrose 4-phosphate. Thus, it can be concluded that it is only the structure of the substrates that prevails in forming a ternary complex enzyme-NAD-thiohemiacetal productive (or not) for hydride transfer in the acylation step. This conclusion is reinforced by the fact that the rate of oxidation for erythrose 4-phosphate by GAPDH is 0.1 s-1 and is limited by the acylation step, whereas glyceraldehyde 3-phosphate acylation is efficient and is not rate-determining (>/=800 s-1). Substituting Asn for His-176 on GapB-encoded protein, a residue postulated to facilitate hydride transfer as a base catalyst, decreases 40-fold the kcat of glyceraldehyde 3-phosphate oxidation. This suggests that the non-efficient positioning of the C-1 atom of glyceraldehyde 3-phosphate relative to the pyridinium of the cofactor within the ternary complex is responsible for the low catalytic efficiency. No phosphorylating activity on erythrose 4-phosphate with GapB-encoded protein is observed although the Pi site is operative as proven by the oxidative phosphorylation of glyceraldehyde 3-phosphate. Thus the binding of inorganic phosphate to the Pi site likely is not productive for attacking efficiently the thioacyl intermediate formed with erythrose 4-phosphate, whereas a water molecule is an efficient nucleophile for the hydrolysis of the thioacyl intermediate. Compared with glyceraldehyde-3-phosphate dehydrogenase activity, this corresponds to an activation of the deacylation step by >/=4.5 kcal.mol-1. Altogether these results suggest subtle structural differences between the active sites of GAPDH and GapB-encoded protein that could be revealed and/or modulated by the structure of the substrate bound. This also indicates that a protein engineering approach could be used to convert a phosphorylating aldehyde dehydrogenase into an efficient non-phosphorylating one and vice versa.  相似文献   

8.
Cell-free extracts of Methanobacterium thermoautotrophicum were found to contain high activities of the following oxidoreductases (at 60 degrees C): pyruvate dehydrogenase (coenzyme A acetylating), 275 nmol/min per mg of protein; alpha-ketoglutarate dehydrogenase (coenzyme A acylating), 100 nmol/min per mg; fumarate reductase, 360 nmol/min per mg; malate dehydrogenase, 240 nmol/min per mg; and glyceraldehyde-3-phosphate dehydrogenase, 100 nmol/min per mg. The kinetic properties (apparent V(max) and K(M) values), pH optimum, temperature dependence of the rate, and specificity for electron acceptors/donors of the different oxidoreductases were examined. Pyruvate dehydrogenase and alpha-ketoglutarate dehydrogenase were shown to be two separate enzymes specific for factor 420 rather than for nicotinamide adenine dinucleotide (NAD), NADP, or ferredoxin as the electron acceptor. Both activities catalyzed the reduction of methyl viologen with the respective alpha-ketoacid and a coenzyme A-dependent exchange between the carboxyl group of the alpha-ketoacid and CO(2). The data indicate that the two enzymes are similar to pyruvate synthase and alpha-ketoglutarate synthase, respectively. Fumarate reductase was found in the soluble cell fraction. This enzyme activity coupled with reduced benzyl viologen as the electron donor, but reduced factor 420, NADH, or NADPH was not effective. The cells did not contain menaquinone, thus excluding this compound as the physiological electron donor for fumarate reduction. NAD was the preferred coenzyme for malate dehydrogenase, whereas NADP was preferred for glyceraldehyde-3-phosphate dehydrogenase. The organism also possessed a factor 420-dependent hydrogenase and a factor 420-linked NADP reductase. The involvement of the described oxidoreductases in cell carbon synthesis is discussed.  相似文献   

9.
pH-titrations with NADH show two ionizable groups in mitochondrial and cytoplasmic malate dehydrogenase, the first with a pKa in the range 6.8-8.3 for the mitochondrial and 6.4-7.8 for the cytoplasmic enzyme, the second with a lower limit at 10.2 resp. 11. Comparison with bis-(dihydronicotinamide)-dinucleotide and dihydronicotinamide-ribosyl-P2-ribose-pyrophosphate instead of NADH indicates that the second alkaline ionization is caused by a residue placed near the adenine binding site of the active centre of the two isoenzymes. Binding studies with NADH and NAD+ give evidence for the participation of a group in the mitochondrial enzyme with pKa 6.8, deprotonation of which is necessary for detectable association of NAD+. In contrast the fixation of NAD+ to the cytoplasmic enzyme is independent of pH.  相似文献   

10.
The mannitol-1-phosphate dehydrogenase (M1PDH) (EC 1.1.1.17) from Streptococcus mutans strain FA-1 was purified to approximately a 425-fold increase in specific activity with a 29% recovery of total enzyme units, using a combination of (i) streptomycin sulfate and ammonium sulfate precipitation and (ii) diethyl-aminoethyl-cellulose (DE-52), agarose A 0.5M, and agarose-nicotinamide adenine dinucleotide (NAD) affinity column chromatography. Polyacrylamide gel electrophoresis of the purified enzyme preparation showed a single protein component that coincided with a band of M1PDH activity. The enzyme had a molecular weight of approximately 45,000 and was stable for long periods of time when stored at -80 degrees C in the presence of beta-mercaptoethanol. Its activity was not affected by mono- or divalent cations, and high concentrations of ethylenedia-minetetraacetic acid were not inhibitory. The M1PDH catalyzed both the NAD-dependent oxidation of mannitol-1-phosphate and the reduced NAD (NADH)-dependent reduction of fructose-6-phosphate. The forward reaction was highly specific for mannitol-1-phosphate and NAD, whereas the reverse reaction was highly specific for NADH and fructose-6-phosphate. The K(m) values for mannitol-1-phosphate and NAD were 0.15 and 0.066 mM, respectively, and the K(m) values for fructose-6-phosphate and NADH were 1.66 and 0.016 mM, respectively. The forward and reverse reactions catalyzed by the M1PDH from S. mutans appeared to be under cellular control. Both adenosine 5'-triphosphate and fructose-6-phosphate were negative effectors of the forward reaction, whereas adenosine 5'-diphosphate served as a negative effector of the reverse reaction catalyzed by the enzyme.  相似文献   

11.
Specialized cells within the aphid, Schizaphis graminum, contain intracellular, vesicle-enclosed eubacterial endosymbionts (Buchnera aphidicola). Using oligonucleotide probes derived from conserved sequences of the ATP synthase beta-subunit and glyceraldehyde-3-phosphate dehydrogenase, and the polymerase chain reaction (PCR), we have amplified, cloned, and sequenced three DNA fragments. Amino acid sequence similarity indicated that two of these fragments corresponded to endosymbiont and host genes potentially coding for the beta-subunit of ATP synthase. The host gene fragment contained two putative introns. The third DNA fragment corresponded to a portion of a gene coding for a glyceraldehyde-3-phosphate dehydrogenase that was highly related to one of the enzymes from Escherichia coli (GapA). These results indicate that B. aphidicola may have an ATP synthase and consequently could synthesize ATP from a proton motive force generated within the intracellular vesicles of host cells containing the endosymbionts. The detection of a gene fragment coding for a protein similar to glyceraldehyde-3-phosphate dehydrogenase suggests the presence of this glycolytic enzyme in the endosymbiont and its involvement in energy-yielding metabolism.  相似文献   

12.
To investigate possible biochemical mechanisms underlying the "toxic gain of function" associated with polyglutamine expansions, the ability of guinea pig liver tissue transglutaminase to catalyze covalent attachments of various polyamines to polyglutamine peptides was examined. Of the polyamines tested, spermine is the most active substrate, followed by spermidine and putrescine. Formation of covalent cross links between polyglutamine peptides and polyamines yields high-M(r) aggregates--a process that is favored with longer polyglutamines. In the presence of tissue transglutaminase, purified glyceraldehyde-3-phosphate dehydrogenase (a key glycolytic enzyme that binds tightly to the polyglutamine domains of both huntingtin and dentatorubral-pallidoluysian atrophy proteins) is covalently attached to polyglutamine peptides in vitro, resulting in the formation of high-M(r) aggregates. In addition, endogenous glyceraldehyde-3-phosphate dehydrogenase of a Balb-c 3T3 fibroblast cell line overexpressing human tissue transglutaminase forms cross-links with a Q60 polypeptide added to the cell homogenate. Possibly, expansion of polyglutamine domains (thus far known to occur in the gene products associated with at least seven neurodegenerative diseases) leads to increased/aberrant tissue transglutaminase-catalyzed cross-linking reactions with both polyamines and susceptible proteins, such as glyceraldehyde-3-phosphate dehydrogenase. Formation of cross-linked heteropolymers may lead to deposition of high-M(r) protein aggregates, thereby contributing to cell death.  相似文献   

13.
Many Oriental people possess a liver mitochondrial aldehyde dehydrogenase where glutamate at position 487 has been replaced by a lysine, and they have very low levels of mitochondrial aldehyde dehydrogenase activity. To investigate the cause of the lack of activity of this aldehyde dehydrogenase, we mutated residue 487 of rat and human liver mitochondrial aldehyde dehydrogenase to a lysine and expressed the mutant and native enzyme forms in Escherichia coli. Both rat and human recombinant aldehyde dehydrogenases showed the same molecular and kinetic properties as the enzyme isolated from liver mitochondria. The E487K mutants were found to be active but possessed altered kinetic properties when compared to the glutamate enzyme. The Km for NAD+ at pH 7.4 increased more than 150-fold, whereas kcat decreased 2-10-fold with respect to the recombinant native enzymes. Detailed steady-state kinetic analysis showed that the binding of NAD+ to the mutant enzyme was impaired, and it could be calculated that this resulted in a decreased nucleophilicity of the active site cysteine residue. The rate-limiting step for the rat E487K mutant was also different from that of the recombinant rat liver aldehyde dehydrogenase in that no pre-steady-state burst of NADH formation was found with the mutant enzyme. Both the rat native enzyme and the E487K mutant oxidized chloroacetaldehyde twice as fast as acetaldehyde, indicating that the rate-limiting step was not hydride transfer or coenzyme dissociation but depended upon nucleophilic attack. Each enzyme form showed a 2-fold activation upon the addition of Mg2+ ions. Substituting a glutamine for the glutamate did not grossly affect the properties of the enzyme. Glutamate 487 may interact directly with the positive nicotinamide ring of NAD+ for the Ki of NADH was the same in the lysine enzyme as it was in the glutamate form. Because of the altered NAD+ binding properties and kcat of the E487K variant, it is assumed that people possessing this form will not have a functional mitochondrial aldehyde dehydrogenase.  相似文献   

14.
The sugar-induced inhibition of malolactic fermentation in cell suspensions of Leuconostoc oenos, recently reclassified as Oenococcus oeni (L. M. T. Dicks, F. Dellaglio, and M. D. Collins, Int. J. Syst. Bacteriol. 45:395-397, 1995) was investigated by in vivo and in vitro nuclear magnetic resonance (NMR) spectroscopy and manometric techniques. At 2 mM, glucose inhibited malolactic fermentation by 50%, and at 5 mM or higher it caused a maximum inhibitory effect of ca. 70%. Galactose, trehalose, maltose, and mannose caused inhibitory effects similar to that observed with glucose, but ribose and 2-deoxyglucose did not affect the rate of malolactic activity. The addition of fructose or citrate completely relieved the glucose-induced inhibition. Glucose was not catabolized by permeabilized cells, and inhibition of malolactic fermentation was not observed under these conditions. 31P NMR analysis of perchloric acid extracts of cells obtained during glucose-malate cometabolism showed high intracellular concentrations of glucose-6-phosphate, 6-phosphogluconate, and glycerol-3-phosphate. Glucose-6-phosphate, 6-phosphogluconate, and NAD(P)H inhibited the malolactic activity in permeabilized cells or cell extracts, whereas NADP+ had no inhibitory effect. The purified malolactic enzyme was strongly inhibited by NADH, whereas all the other above-mentioned metabolites exerted no inhibitory effect, showing that NADH was responsible for the inhibition of malolactic activity in vivo. The concentration of NADH required to inhibit the activity of the malolactic enzyme by 50% was ca. 25 microM. The data provide a coherent biochemical basis to understand the glucose-induced inhibition of malolactic fermentation in L. oenos.  相似文献   

15.
Three homozygous allelic forms (alpha GPDHf, alpha GPDHm and alpha GPDHs) of NAD+-dependent glycerol-3-phosphate dehydrogenase (sn-glycerol-3-phosphate:NAD+ 2-oxidoreductase, EC 1.1.1.8) of Drosophila virilis were purified to homogeneity and their biochemical properties were compared. Although these three forms were mutually distinguishable by electrophoresis, no significant differences were found with respect to pH optima for both forward and reverse reactions (pH 6.0--6.5 for dihydroxyacetone phosphate reduction; pH 10.0--10.5 for glycerol 3-phosphate oxidation), native and subunit molecular weights (65 000 for native form; 35 000--37 000 for subunit) and Michaelis constants for NADH, glycerol 3-phosphate and NAD+ (5.3--6.0 microM for NADH; 1.8--1.9 mM for glycerol 3-phosphate; 100--110 microM for NAD+). Significant differences among three forms were observed in thermostability at 35 degrees C and inhibition by excess of dihydroxyacetone phosphate. The alpha GPDHf form was found to be most thermolabile and the alpha GPDHs form most susceptible to the inhibition.  相似文献   

16.
In primary cocultures of neurons and glial cells prepared from the neonatal rat brain, lipopolysaccharide (LPS) reduced the numbers of neuronal cells but the effects were markedly inhibited by NG-monomethyl-L-arginine, indicating the involvement of NO and LPS-induced NO synthase in neuronal death. LPS stimulated the expression of inducible NOS (iNOS) in preparations of primary cultured microglias/astrocytes, but not in primary cultured neurons. In addition, LPS caused DNA fragmentation only in NG108-15 cells but not in primary cultured astrocytes as well as astrocytes in cocultures of the two cell types, suggesting that NOS induces the apoptosis of neurons but not glial cells. We then examined the NO-induced neuronal death in NG108-15 cells using NO donors. SNP, and NO donor, caused NO-2 accumulation in the reaction medium and lactate dehydrogenase (LDH) leakage from NG108-15 cells. Although SNP stimulated guanylyl cyclase and accumulated cGMP, cGMP analogs did not affect LDH leakage. In addition, SNP induced chromosomal condensation and fragmentation of nuclei in NG108-15 cells. Gel electrophoretic analysis of cellular DNA extracted from SNP-treated cells, confirmed the internucleosomal DNA fragmentation typical of apoptosis in this culture. SNP increased the amount of radioisotopic labeled glyceraldehyde-3 phosphate dehydrogenase (GAPDH) in the presence of [32P]NAD and inhibited the enzyme activity. The results suggested that SNP-induced cell death is partly due to the NO-induced inhibition of GAPDH, perhaps by stimulating the binding of NAD to GAPDH.  相似文献   

17.
BACKGROUND: The leucine biosynthetic enzyme 3-isopropylmalate dehydrogenase (IMDH) belongs to a unique class of bifunctional decarboxylating dehydrogenases. The two best-known members of this family, IMDH and isocitrate dehydrogenase (IDH), share a common structural framework and catalytic mechanism but have different substrate and cofactor specificities. IMDH is NAD(+)-dependent, while IDHs occur in both NAD(+)-dependent and NADP(+)-dependent forms. RESULTS: We have co-crystallized Thermus thermophilus IMDH with NAD+ and have determined the structure at 2.5 A resolution. NAD+ binds in an extended conformation. Comparisons with the structure in the absence of cofactor show that binding induces structural changes of up to 2.5 A in the five loops which form the dinucleotide-binding site. The adenine and diphosphate moieties of NAD+ are bound via interactions which are also present in the NADP(+)-IDH complex. Amino acids which interact with the NADP+ 2'-phosphate in IDH are substituted or absent in IMDH. The adenosine ribose forms two hydrogen bonds with Asp278, and the nicotinamide and nicotinamide ribose interact with Glu87 and Asp78, all unique to IMDH. CONCLUSIONS: NAD+ binding induces a conformational transition in IMDH, resulting in a structure that is intermediate between the most 'open' and 'closed' decarboxylating dehydrogenase conformations. Physiological specificity of IMDH for NAD+ versus NADP+ can be explained by the unique interaction between Asp278 and the free 2'-hydroxyl of the NAD+ adenosine, discrimination against the presence of the 2'-phosphate by the negative charge on Asp278, and the absence of potential favorable interactions with the 2'-phosphate of NADP+.  相似文献   

18.
Inosine 5'-monophosphate dehydrogenase (IMPDH) catalyzes the oxidation of inosine 5-monophosphate (IMP) to xanthosine 5'-monophosphate (XMP). The reaction proceeds with concomitant conversion of NAD+ to NADH and is the rate-limiting step in the de novo biosynthesis of guanosine nucleotides. IMPDH is a target for numerous chemotherapeutic agents. The conformations of enzyme-bound substrates, enzyme-bound products and enzyme-bound ligands in general, are of interest for the understanding of the catalytic mechanism of the enzyme and the design of new inhibitors. Although several of the chemotherapeutic inhibitors of IMPDH are NAD+ or NADH analogues, no structural data for IMPDH-bound NAD+ (or NADH) are available. In the present work, we have used transferred nuclear Overhauser effect spectroscopy (TRNOESY) to determine the conformation of NADH bound to the active site of human type II IMPDH (IMPDH-h2). The inter-proton distances determined from TRNOESY data indicate that NADH binds to the enzyme active site in an overall extended conformation. The adenosine moiety and the nicotinamide riboside moiety are both in the anti conformation about the glycosidic bond, and both ribose rings are in approximately C4'-exo conformations. The nicotinamide amide group was found to be in a cis conformation. The anti conformation of the nicotinamide riboside moiety is in accord with the preferred conformations of several potent and selective dinucleotide inhibitors and is consistent with that implied by the stereospecificity of hydride transfer in the enzymatic reaction. The implications of this conformation for the catalytic mechanism of IMPDH-h2 are discussed.  相似文献   

19.
The tetrazolium salt 5-cyano-2,3-di-p-toluyl-tetrazolium chloride (CTC), yielding a fluorescent formazan on reduction, was used to measure NAD(P)H oxidoreductase activity. In this study, optimal conditions for the flow cytometric technique were determined empirically with tissue culture cell lines and mouse Ehrlich ascites cells. Applying a coupled reaction procedure, NADH and NADPH as substrates of the oxidoreductases to be measured are generated endogenously by lactate or glucose-6-phosphate dehydrogenase, respectively. The results were evaluated by combining spectrophotometry and flow cytometry. We obtained integral activities for each group of NADH and NADPH oxidoreductases. Furthermore, by counterstaining the DNA with DAPI, followed by bivariate analysis of flow cytometric data, our assay gives a detailed distribution of enzyme activities of all cells, even in subgroups present in heterogeneous cell populations. Therefore, this protocol permits the study of NAD(P)H oxidoreductase activities in ex vivo tumor samples in which mixed cellular populations may be present.  相似文献   

20.
The periplasmic nucleotide pyrophosphatase from Haemophilus parasuis was purified 750-fold to electrophoretic homogeneity through salt fractionation and ion-exchange and affinity chromatography. The purified enzyme was monomeric with an apparent M(r) of 70,000 and catalyzed the hydrolysis of the pyrophosphate bond of NAD to yield NMN and AMP as products. The enzyme exhibited negative cooperativity in the hydrolysis of a number of pyridine dinucleotides and structurally-related pyrophosphate compounds as indicated by biphasic double-reciprocal plots and Hill coefficients of 0.5. The kinetic parameters, K(m) and Vm, determined titrimetrically and analyzed through computer programs, were used to compare the relative effectiveness of dinucleotides containing nitrogen bases other than nicotinamide or adenine to that of NAD. Effective substrate-competitive inhibition of the pyrophosphatase was observed with purine and pyrimidine nucleoside diphosphates in the low micromolar concentration range. Although less effective, N1-alkylnicotinamide chlorides also inhibited competitively with respect to the substrate, NAD. In addition to being an effective inhibitor of the purified enzyme, adenosine diphosphate also inhibited growth of H. parasuis at a low micromolar concentration. This inhibition of growth correlates well with inhibition of the periplasmic pyrophosphatase which is supported by the fact that adenosine diphosphate does not effectively inhibit growth when the pyrophosphatase is by-passed by growth on nicotinamide mononucleotide. These observations are all consistent with the periplasmic nucleotide pyrophosphatase being essential for the growth of the organism on NAD and therefore, a very important enzyme with respect to the pathogenesis of the organism. 3-Aminopyridine mononucleotide, which also inhibited growth of H. parasuis at a low micromolar concentration, did not effectively inhibit the purified pyrophosphatase and a different target enzyme needs to be considered to explain growth inhibition by this derivative.  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