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 406 毫秒
1.
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   

2.
The kinetics of the NADP+- and phosphate-dependent oxidation of glutamic acid 5-semialdehyde are consistent with a rapid-equilibrium random order mechanism. The Km for DL-pyrroline-5-carboxylic acid is 2.5 mM, for NADP+ is 0.05 mM and for phosphate is 0.35 mM. The Vmax is approx. 8.0 units per mg protein. The reaction is highly specific for the DL-pyrroline-5-carboxylic acid and NADP+, but a number of divalent anions can substitute for phosphate. NADPH is competitive with respect to all three substrates and an analog of gamma-glutamyl phosphate, 3-(phosphonoacetylamido)-L-alanine, is competitive with respect to DL-pyrroline-5-carboxylic acid and non-competitive with respect to NADP+ and phosphate, suggesting dead-end complex formation.  相似文献   

3.
NADP-dependent isocitrate dehydrogenase activity has been screened in several cyanobacteria grown on different nitrogen sources; in all the strains tested isocitrate dehydrogenase activity levels were similar in cells grown either on ammonium or nitrate. The enzyme from the unicellular cyanobacterium Synechocystis sp. PCC 6803 has been purified to electrophoretic homogeneity by a procedure that includes Reactive-Red-120-agarose affinity chromatography and phenyl-Sepharose chromatography as main steps. The enzyme was purified about 600-fold, with a yield of 38% and a specific activity of 15.7 U/mg protein. The native enzyme (108 kDa) is composed of two identical subunits with an apparent molecular mass of 57 kDa. Synechocystis isocitrate dehydrogenase was absolutely specific for NADP as electron acceptor. Apparent Km values were 125, 59 and 12 microM for Mg2+, D,L-isocitrate and NADP, respectively, using Mg2+ as divalent cation and 4, 5.7 and 6 microM for Mn2+, D,L-isocitrate and NADP, respectively, using Mn2+ as a cofactor. The enzyme was inhibited non-competitively by ADP (Ki, 6.4 mM) and 2-oxoglutarate, (Ki, 6 mM) with respect to isocitrate and in a competitive manner by NADPH (Ki, 0.6 mM). The circular-dichroism spectrum showed a protein with a secondary structure consisting of about 30% alpha-helix and 36% beta-pleated sheet. The enzyme is an acidic protein with an isoelectric point of 4.4 and analysis of the NH2-terminal sequence revealed 45% identity with the same region of Escherichia coli isocitrate dehydrogenase. The aforementioned data indicate that NADP isocitrate dehydrogenase from Synechocystis resembles isocitrate dehydrogenase from prokaryotes and shows similar molecular and structural properties to the well-known E. coli enzyme.  相似文献   

4.
Electrophoretic studies were performed on enzymes concerned with the oxidation of malate in free-living and bacteroid cells of Mesorhizobium ciceri CC 1192, which forms nitrogen-fixing symbioses with chickpea (Cicer arietinum L.) plants. Two malate dehydrogenases were detected in extracts from both types of cells in native polyacrylamide electrophoresis gels that were stained for enzyme activity. One band of malate dehydrogenase activity was stained only in the presence of NADP+, whereas the other band was revealed with NAD+ but not NADP+. Further evidence for the occurrence of separate NAD- and NADP-dependent malate dehydrogenases was obtained from preliminary enzyme kinetic studies with crude extracts from free-living M. ciceri CC 1192 cells. Activity staining of electrophoretic gels also indicated the presence of two malic enzymes in free-living and bacteroid cells of M. ciceri CC 1192. One malic enzyme was active with both NAD+ and NADP+, whereas the other was specific for NADP+. Possible roles of the multiple forms of malate dehydrogenase and malic enzyme in nitrogen-fixing symbioses are discussed.  相似文献   

5.
The chemical shifts of all the aromatic proton and anomeric proton resonances of NADP+, NADPH, and several structural analogues have been determined in their complexes with Lactobacillus casei dihydrofolate reductase by double-resonance (saturation transfer) experiments. The binding of NADP+ to the enzyme leads to large (0.9-1.6 ppm) downfield shifts of all the nicotinamide proton resonances and somewhat smaller upfield shifts of the adenine proton resonance. The latter signals show very similar chemical shifts in the binary and ternary complexes of NADP+ and the binary complexes of several other coenzymes, suggesting that the environment of the adenine ring is similar in all cases. In contrast, the nicotinamide proton resonances show much greater variability in position from one complex to another. The data show that the environments of the nicotinamide rings of NADP+, NADPH, and the thionicotinamide and acetylpyridine analogues of NADP+ in their binary complexes with the enzyme are quite markedly different from one another. Addition of folate or methotrexate to the binary complex has only modest effects on the nicotinamide ring of NADP+, but trimethoprim produces a substantial change in its environment. The dissociation rate constant of NADP+ from a number of complexes was also determined by saturation transfer.  相似文献   

6.
Reaction of glycolaldehyde with the binary E-NADP complex of bovine kidney aldose reductase (ALR2) produces an enzyme-bound chromophore whose absorbance (lambd max 341 nm) and fluorescence (lambda ex max 341 nm; lambda emit max 421 nm) properties are distinct from those of NADPH or E.NADPH yet are consistent with the proposed covalent adduct structure [1,4-dihydro-4-(1-hydroxy-2-oxoethyl)nicotinamide adenine dinucleotide phosphate]. The kinetics of adduct formation, both in solution and at the enzyme active site, support a mechanism involving rate-determining enolization of glycolaldehyde at high [NADP+] or [E.NADP]. At low [NADP+] or [E.NADP] the reaction is second-order overall, but the ALR2-mediated reaction displays saturation by glycolaldehyde due to competition of the aldehyde (plus hydrate) and enol for E.NADP. Measurement of the pre-steady-state burst of E-adduct formation confirms that glycolaldehyde enol is the reactive species and gives a value of 1.3 x 10(-6) for Kenol = [enol]/[( aldehyde] + [hydrate]), similar to that determined by trapping the enol with I3-. At the ALR2 active site, the rate of adduct formation is enhanced 79,000-fold and the adduct is stabilized greater than or equal to 13,000-fold relative to the reaction with NADP+ in solution. A portion of this enhancement is ascribed to specific interaction of NADP+ with the enzyme since the 3-acetylpyridine analogue, (AP)ADP+, gives values that are 15-200-fold lower. Additional evidence for strong interaction of ALR2 with both NADP+ and NADPH is reported. Yet, because dissociation of adduct is slow, catalysis of the overall adduct formation reaction by ALR2 is less than or equal to 67-fold.  相似文献   

7.
The interactions of calcium with NAD+, NADH, NADP+ and NADPH in a 50% (by volume) methanol/water mixture (pH 7, 25 degrees C) were studied by calorimetry. The association constants for 1:1 complex formation were found to be 6.6 +/- 0.2, 270 +/- 76, 18 +/- 3 and 98 +/- 10 for NAD+, NADH, NADP+ and NADPH, respectively. Comparing these to the association constants for an aqueous system reveals that as the polarity of the solvent system is decreased the interactions involving NAD+, NADP+ and NADPH are all decreased. In contrast, the interaction involving NADH is markedly increased. All the interactions were found to be endothermic.  相似文献   

8.
A simple and rapid method is presented for purifying the NADP+-dependent isocitrate dehydrogenase (threo-DS-isocitrate:NADP+ oxidoreductase (decarboxylating), from Escherichia coli, which relies on fractionation of the enzyme with polyethylene glycol. The shortened preparation results in a 32% relative recovery of purified enzyme at a specific activity of 127 micronmol/min per mg of protein. The Km values for threo-DS-isocitrate, NADP+, NAD+, Mg2+ and Mn2+ are 6.4, 36, 3000, 19.7 and 2.0 micronM, respectively. The stability of the enzyme as a function of dilution and temperature are also reported. Recrystallization of the purified enzyme under different conditions readily produces a variety of single crystals. Crystals grown from ammonium sulfate solutions belong to monoclinic space group C2 with a = 125 A, b = 111 A, c = 83.5 A and beta = 108degrees 45'. Density measurements of these crystals indicate there are two 80 000-dalton dimers per asymmetric unit.  相似文献   

9.
The kinetic properties for the native forward reaction of pyruvate:NADP+ oxidoreductase from Euglena gracilis were determined. The substrate kinetics gave a pattern of a ping-pong mechanism involving a competitive substrate inhibition of CoA against pyruvate. The Km values for pyruvate, CoA, and NADP+ were estimated to be 27, 6.6, and 28 microM, respectively, and the Ki value of CoA against pyruvate was 28 microM. CO2 inhibited noncompetitively against pyruvate and NADP+, and uncompetitively against CoA. Acetyl-CoA showed a competitive inhibition with respect to pyruvate and an uncompetitive inhibition with respect to NADP+. NADPH inhibited competitively versus NADP+, noncompetitively versus CoA, and uncompetitively versus pyruvate. The kinetic behavior is consistent with a two-site ping-pong mechanism involving the substrate inhibition. From the kinetic mechanism, it is proposed that the enzyme has two catalytic sites linked by an intramolecular electron-transport chain. One of these is a thiamine pyrophosphate-containing catalytic site which reacts with pyruvate and CoA to form CO2 and acetyl-CoA, and the other site functions in the reduction of NADP+. In contrast, when methyl viologen was used as an artificial one-electron acceptor substituting for NADP+, the reaction gave a pattern characteristic of an octa uni ping-pong mechanism involving a competitive substrate inhibition of CoA against pyruvate.  相似文献   

10.
1. 6-phosphogluconate dehydrogenase from sheep liver has been purified 350-fold by affinity chromatography with a final specific activity of 18 micronmol of NADP+/reduced min per mg of protein and an overall yield of greater than 40%. 2. A systematic investigation of potential ligands has been carried out: these included 6-phosphogluconate and NADP+, pyridoxal phosphate and several immobilized nucleotides. The results indicate that NADP+ is the most suitable ligand for the purification of 6-phosphogluconate dehydrogenase. 3. The effects of pH and alternative eluents have been examined in relation to the parameters known to affect the desorption phase of affinity chromatography; careful manipulation of the elution conditions permitted the separation of glucose 6-phosphate dehydrogenase, glutathione reductase and 6-phosphogluconate dehydrogenase from sheep liver on NADP+-Sepharose 4B. 4. A large-scale purification scheme for 6-phosphogluconate dehydrogenase is presented that uses the competitive inhibitors inorganic pyrophosphate and citrate as specific eluents.  相似文献   

11.
Nitrogen starvation enhances up to 8-fold the cellular level of the NADP+-dependent isocitrate dehydrogenase activity (isocitrate:NADP+ oxidoreductase (decarboxylating), IDH, EC 1.1.1.42) in the thermophilic filamentous non-N2-fixing cyanobacterium Phormidium laminosum. The enzyme was purified 650-fold to electrophoretic homogeneity from nitrogen-starved cells with an activity yield of 25% and a specific activity of 500 U (mg protein)-1. The native enzyme showed a pI of 5.9 and it was a dimer of 107 kDa consisting of two identical subunits of 53 kDa. The activity required the presence of a divalent metal cation as an essential activator, Mn2+ or Mg2+ being the most effective. The optimum temperature for activity was 55 degrees C and the Ea for catalysis was 39.7 kJ mol-1. An optimum pH for activity of 8.5 was found and the calculated pKE1, pKE2 and pKES1 of enzyme ionisation groups were 6.0, 8.9 and 6.3, respectively. Km values of 22, 50 and 24 microM were calculated for d,l-isocitrate, NADP and Mn2+, respectively, in the Mn2+-dependent reaction and 70, 32 and 159 microM for d,l-isocitrate, NADP and Mg2+, respectively, in the Mg2+-dependent reaction. The decarboxylating activity was inhibited by ATP, ADP and by its reaction products 2-oxoglutarate and NADPH2. Polyclonal antibodies raised against the pure IDH were used to assess the presence of the enzyme in cells subjected to nitrogen starvation.  相似文献   

12.
1. The stoicheiometries and affinities of ligand binding to isocitrate dehydrogenase were studied at pH 7.0, mainly by measuring changes in NADPH and protein fluorescence. 2. The affinity of the enzyme for NADPH is about 100-fold greater than it is for NADP+ in various buffer/salt solutions, and the affinities for both coenzymes are decreased by Mg2+, phosphate and increase in ionic strength. 3. The maximum binding capacity of the dimeric enzyme for NADPH, from coenzyme fluorescence and protein-fluorescence measurements, and also for NADP+, by ultrafiltration, is 2 mol/mol of enzyme. Protein-fluorescence titrations of the enzyme with NADP+ are apparently inconsistent with this conclusion, indicating that the increase in protein fluorescence caused by NADP+ binding is not proportional to fractional saturation of the binding sites. 4. Changes in protein fluorescence caused by changes in ionic strength and by the binding of substrates, Mg2+ or NADP+ (but not NADPH) are relatively slow, suggesting conformation changes. 5. In the presence of Mg2+, the enzyme binds isocitrate very strongly, and 2-oxoglutarate rather weakly. 6. Evidence is presented for the formation of an abortive complex of enzyme-Mg2+-isocitrate-NADPH in which isocitrate and NADPH are bound much more weakly than in their complexes with enzyme and Mg2+ alone. 7. The results are discussed in relation to the interpretation of the kinetic properties of the enzyme and its behaviour in the mitochondrion.  相似文献   

13.
Glucose-6-phosphate dehydrogenase from sporangiophores of Phycomyces blakesleeanus NRRL 1555 (-) was partially purified. The enzyme showed a molecular weight of 85 700 as determined by gel-filtration. NADP+ protected the enzyme from inactivation. Magnesium ions did not affect the enzyme activity. Glucose-6-phosphate dehydrogenase was specific for NADP+ as coenzyme. The reaction rates were hyperbolic functions of substrate and coenzyme concentrations. The Km values for NADP+ and glucose 6-phosphate were 39.8 and 154.4 microM, respectively. The kinetic patterns, with respect to coenzyme and substrate, indicated a sequential mechanism. NADPH was a competitive inhibitor with respect to NADP+ (Ki = 45.5 microM) and a non-competitive inhibitor with respect to glucose 6-phosphate. ATP inhibited the activity of glucose-6-phosphate dehydrogenase. The inhibition was of the linear-mixed type with respect to NADP+, the dissociation constant of the enzyme-ATP complex being 2.6 mM, and the enzyme-NADP+-ATP dissociation constant 12.8 mM.  相似文献   

14.
The binding site of NADPH in NADPH-adrenodoxin reductase was examined using crystalline enzyme from bovine adrenocortical mitochondria by studies on the effects of photooxidation and chemical modifications of amino acid residues in the reductase. (1) Photoxication decreased the enzymatic activity of NADPH-adrenodoxin reductase. Photooxidation of the reductase was prevented by NADP+, adrenodoxin, or reduced glutathione, but not NAD+. Photoinactivation caused loss of a histidyl residue, but not of tyrosyl, tryptophanyl, cysteinyl, or methionyl residues of the reductase. It did not affect the circular dichroism spectrum of the reductase appreciably. (2) NADPH-adrenodoxin reductase activity was inhibited by diethyl pyrocarbonate and the inhibition was partially reversed by addition of hydroxylamine. The inhibition was prevented by NADP+, but not NAD+. (3) NADPH-adrenodoxin reductase activity was inhibited by 5,5'-dithiobis(2-nitrobenzoate) and the inhibition was reversed by reduced glutathione. It was also protected by NADP+, but not NAD+. The results indicate that a histidyl residue and a cysteinyl residue of NADPH-adrenodoxin reductase are essential for the binding of NADPH by the reductase.  相似文献   

15.
NADP+, NADPH and glucose 6-phosphate dehydrogenase were determined in the cerebral cortex of mice exposed to high O2 pressure for 0, 8 and 16 min. These time intervals corresponded to 0, 50 and 100% of the CT50 (the time taken for 50% of the mice to convulse). Cerebral NADP+, NADPH and glucose 6-phosphate dehydrogenase also were determined in O2-exposed mice exhibiting hyperactivity, convulsions, and in mice killed 10s after convulsions. Similar increases in cortical NADP+ and decreases in NADPH were found in mice exposed to 610kPa (6 atm.) of 100% O2 for 0, 50 and 100% of the CT50, during hyperactivity, onset of seizure and 10s after convulsions. The NADP+/NADPH ratio increased approx. 25% at 0% of the CT50, and remained at this increased value at all O2-exposure periods including the hyperactive state, onset of seizure and 10s after convulsions. Identical changes in cerebral NADP+ , NADPH and the NADP+/NADPH ratio were found in mice exposed for 16min to 100% O2 at 100, 350 or 610kPa. No change in cerebral glucose 6-phosphate dehydrogenase was found in mice exposed to 610kPa of 100% O2 during the various stages of O2 toxicity. Only in the 10s post-convulsive group was a statistically significant decrease in glucose 6-phosphate dehydrogenase observed. Disulfiram [bis(diethylthiocarbamoyl) disulphide], an effective O2-protective agent, did not prevent the O2-induced increase in cerebral NADP+ and the NADP+/NADPH ratio, or decrease in NADPH.  相似文献   

16.
辅酶NAD(H)相比NADP(H)有稳定性好、价格低廉及更广的辅酶循环方法等优势,因此在实际应用中常需将NADP(H)依赖型的脱氢酶改造成为NAD(H)依赖型的。来源于嗜热共生杆菌Symbiobacterium thermophilum的NADP(H)依赖型内消旋-2,6-二氨基庚二酸脱氢酶(meso-2,6-diaminopimelate dehydrogenase,St DAPDH)及其突变体酶是催化还原氨化合成D-氨基酸的优良催化剂,本研究试图改变其辅酶偏好性,增强其应用优势。对其晶体结构分析可知,氨基酸残基Y76距离腺嘌呤较近,R35及R36和辅酶上磷酸基团有直接相互作用。依氨基酸侧链基团性质对Y76进行了定点突变,发现不同突变子对两种辅酶的偏好性都发生了变化;对与磷酸基团直接作用的R35、R36进行的双突变R35S/R36V,导致酶对NADP+的催化活力降低;将R35S/R36V和部分Y76突变进行了组合,发现三突变组合以NAD+为辅酶时的活力均大于以NADP+为辅酶的活力,实现了辅酶偏好性转变。这些研究工作为进一步实现St DAPDH的辅酶偏好性完全转变提供依据。  相似文献   

17.
ADP-ribosyl cyclases (ADPRCs) are present from lower Metazoa to mammals and synthesize the Ca2+-active (di)nucleotides cyclic ADP-ribose (cADPR), NAADP+, and ADP-ribose (ADPR), involved in the regulation of important cellular functions. NAADP+ can be synthesized by ADPRCs from NADP+ through a base-exchange reaction, which substitutes nicotinamide for nicotinic acid (NA). Here we demonstrate that ADPRCs from both lower and higher Metazoa (including human CD38) can also synthesize NAADP+ starting from 2'-phospho-cyclic ADP-ribose (cADPRP) and NA. Comparison, on the two substrates cADPRP and NADP+, of the relative rates of the reactions introducing NA and hydrolyzing/cyclizing the substrate, respectively, indicates that with all ADPRCs tested cADPRP is preferentially transformed into NAADP+, while NADP+ is preferentially cyclized or hydrolyzed to cADPRP/2'-phospho-ADP-ribose. cADPRP was detectable in retinoic acid-differentiated, CD38+ HL-60 cells, but not in undifferentiated, CD38- cells. These results suggest that cADPRP may be a NAADP+ precursor in ADPRC+ cells.  相似文献   

18.
The effect of NADP+ on light-induced steady-state redox changes of membrane-bound cytochromes was investigated in membrane fragements prepared from the blue-green algae Nostoc muscorum (Strain 7119) that had high rates of electron transport from water to NADP+ and from an artificial electron donor, reduced dichlorophenolindophenol (DCIPH2) to NDAP+. The membrane fragments contained very little phycocyanin and had excellent optical properties for spectrophotometric assays. With DCIPH2 as the electron donor, NADP+ had no effect on the light-induced redox changes of cytochromes: with or without NADP+, 715- or 664-nm illumination resulted mainly in the oxidation of cytochrome f and of other component(s) which may include a c-type cytochrome with an alpha peak at 549nm. With 664 nm illumination and water as the electron donor, NADP+ had a pronounced effect on the redox state of cytochromes, causing a shift toward oxidation of a component with a peak at 549 nm (possibly a c-type cytochrome), cytochrome f, and particularly cytochrome b559. Cytochrome b559 appeared to be a component of the main noncyclic electron transport chain and was photooxidized at physiological temperatures by Photosystem II. This photooxidation was apparent only in the presence of a terminal acceptor (NADP+) for the electron flow from water.  相似文献   

19.
Shikimate 5-dehydrogenase (SKDH; EC 1.1.1.25) catalyzes the reversible reduction of 3-dehydroshikimate to shikimate and is a key enzyme in the aromatic amino acid biosynthesis pathway. The shikimate 5-dehydrogenase gene, aroE, from Archaeoglobus fulgidus was cloned and overexpressed in Escherichia coli. The recombinant enzyme purified as a homodimer and yielded a maximum specific activity of 732 U/mg at 87 degrees C (with NADP+ as coenzyme). Apparent Km values for shikimate, NADP+, and NAD+ were estimated at 0.17+/-0.03 mM, 0.19+/-0.01 mM, and 11.4+/-0.4 mM, respectively. The half-life of the A. fulgidus SKDH is 2 h at the assay temperature (87 degrees C) and 17 days at 60 degrees C. Addition of 1 M NaCl or KCl stabilized the enzyme's half-life to approximately 70 h at 87 degrees C and approximately 50 days at 60 degrees C. This work presents the first kinetic analysis of an archaeal SKDH.  相似文献   

20.
The yeast Candida parapsilosis possesses two routes of electron transfer from exogenous NAD(P)H to oxygen. Electrons are transferred either to the classical cytochrome pathway at the level of ubiquinone through an NAD(P)H dehydrogenase, or to an alternative pathway at the level of cytochrome c through another NAD(P)H dehydrogenase which is insensitive to antimycin A. Analyses of mitoplasts obtained by digitonin/osmotic shock treatment of mitochondria purified on a sucrose gradient indicated that the NADH and NADPH dehydrogenases serving the alternative route were located on the mitochondrial inner membrane. The dehydrogenases could be differentiated by their pH optima and their sensitivity to amytal, butanedione and mersalyl. No transhydrogenase activity occurred between the dehydrogenases, although NADH oxidation was inhibited by NADP+ and butanedione. Studies of the effect of NADP+ on NADH oxidation showed that the NADH:ubiquinone oxidoreductase had Michaelis-Menten kinetics and was inhibited by NADP+, whereas the alternative NADH dehydrogenase had allosteric properties (NADH is a negative effector and is displaced from its regulatory site by NAD+ or NADP+).  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