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 905 毫秒
1.
Summary The redox interconversion of Escherichia coli glutathione reductase has been studied both in situ, with permeabilized cells treated with different reductants, and in vivo, with intact cells incubated with compounds known to alter their intracellular redox state.The enzyme from toulene-permeabilized cells was inactivated in situ by NADPH, NADH, dithionite, dithiothreitol, or GSH. The enzyme remained, however, fully active upon incubation with the oxidized forms of such compounds. The inactivation was time-, temperature-, and concentration-dependent; a 50% inactivation was promoted by just 2 M NADPH, while 700 M NADH was required for a similar effect. The enzyme from permeabilized cells was completely protected against redox inactivation by GSSG, and to a lesser extent by dithiothreitol, GSH, and NAD(P)+. The inactive enzyme was efficiently reactivated in situ by physiological GSSG concentrations. A significant reactivation was promoted also by GSH, although at concentrations two orders of magnitude below its physiological concentrations. The glutathione reductase from intact E. coli cells was inactivated in vivo by incubation with DL-malate, DL-isocitrate, or higher L-lactate concentrations. The enzyme was protected against redox inactivation and fully reactivated by diamide in a concentration-dependent fashion. Diamide reactivation was not dependent on the synthesis of new protein, thus suggesting that the effect was really a true reactivation and not due to de novo synthesis of active enzyme. The glutathione reductase activity increased significantly after incubation of intact cells with tert-butyl or cumene hydroperoxides, suggesting that the enzyme was partially inactive within such cells. In conclusion, the above results show that both in situ and in vivo the glutathione reductase of Escherichia coli is subjected to a redox interconversion mechanism probably controlled by the intracellular NADPH and GSSG concentrations.  相似文献   

2.
Summary The glutathione reductase from E. coli was rapidly inactivated following aerobic incubation of the pure and cell-free extract enzymes with NADPH, NADH and other reductants. The inactivation of the pure enzyme depended on the time and temperature of incubation (t1/2 = 2 min at 37°C), and was proportional to the |INADPH|/|enzyme| ratio, reaching 50% in the presence of 0.3 M NADPH and 45 M NADH respectively, at a subunit concentration of 20 nM. Higher pyridine nucleotide concentrations were required to inactivate the enzyme from cell-free extracts. Two apparent pKa, corresponding to pH 5.8 and 7.3, were determined for the redox inactivation. The enzyme remained inactive even after eliminating the excess NADPH by gel chromatography. E. coli glutathione reductase was protected by oxidized and reduced glutathione against redox inactivation with both pure and cell-free extract enzymes. Ferricyanide and dithiothreitol protected only the pure enzyme, while NADP+ exclusively protected the cell-free extract enzyme. The inactive glutathione reductase was reactivated by treatment with oxidized and reduced glutathione, ferricyanide, and dithiothreitol in a time-and temperature-dependent process. The oxidized form of glutathione was more efficient and specific than the reduced form in the protection and reactivation of the pure enzyme.The molecular weight of the redox-inactivated E. coli glutathione reductase was similar to that of the dimeric native enzyme, ruling out aggregation as a possible cause of inactivation. A tentative model is discussed for the redox inactivation, involving the formation of an erroneous disulfide bridge at the glutathione-binding site.  相似文献   

3.
Summary Redox inactivation of glutathione reductase involves metal cations, since chelators protected against NADPH-inactivation, 3 µM EDTA or 10 µM DETAPAC yielding full protection. Ag+, Zn2+ and Cd2+ potentiated the redox inactivation promoted by NADPH alone, while Cr3+, Fe2+, Fe3+, Cu+, and Cu2+ protected the enzyme. The Zn2+ and Cd2+ effect was time-dependent, unlike conventional inhibition. Glutathione reductase interconversion did not require dioxygen, excluding participation of active oxygen species produced by NADPH and metal cations. One Zn2+ ion was required per enzyme subunit to yield full NADPH-inactivation, the enzyme being reactivated by EDTA. Redox inactivation of glutathione reductase could arise from the blocking of the dithiol formed at the active site of the reduced enzyme by metal cations, like Zn2+ or Cd2+.The glutathione reductase activity of yeast cell-free extracts was rapidly inactivated by low NADPH or moderate NADH concentrations; NADP+ also promoted rapid inactivation in fresh extracts, probably after reduction to NADPH. Full inactivation was obtained in cell-free extracts incubated with glucose-6-phosphate or 6-phosphogluconate; the inactivating efficiency of several oxidizable substrates was directly proportional to the specific activities of the corresponding dehydrogenases, confirming that redox inactivation derives from NADPH formed in vitro.Abbreviations DETAPAC diethylenetriaminepentaacetic acid - 2,5-ADP-Sepharose-N6-(6-aminohexyl) adenosine 2,5-bisphosphateSepharose  相似文献   

4.
Glutathione reductase (EC 1.6.4.2) was purified from spinach (Spinacia oleracea L.) leaves by affinity chromatography on ADP-Sepharose. The purified enzyme has a specific activity of 246 enzyme units/mg protein and is homogeneous by the criterion of polyacrylamide gel electrophoresis on native and SDS-gels. The enzyme has a molecular weight of 145,000 and consists of two subunits of similar size. The pH optimum of spinach glutathione reductase is 8.5–9.0, which is related to the function it performs in the chloroplast stroma. It is specific for oxidised glutathione (GSSG) but shows a low activity with NADH as electron donor. The pH optimum for NADH-dependent GSSG reduction is lower than that for NADPH-dependent reduction. The enzyme has a low affinity for reduced glutathione (GSH) and for NADP+, but GSH-dependent NADP+ reduction is stimulated by addition of dithiothreitol. Spinach glutathione reductase is inhibited on incubation with reagents that react with thiol groups, or with heavymetal ions such as Zn2+. GSSG protects the enzyme against inhibition but NADPH does not. Pre-incubation of the enzyme with NADPH decreases its activity, so kinetic studies were performed in which the reaction was initiated by adding NADPH or enzyme. The Km for GSSG was approximately 200 M and that for NADPH was about 3 M. NADP+ inhibited the enzyme, assayed in the direction of GSSG reduction, competitively with respect to NADPH and non-competitively with respect to GSSG. In contrast, GSH inhibited non-competitively with respect to both NADPH and GSSG. Illuminated chloroplasts, or chloroplasts kept in the dark, contain equal activities of glutathione reductase. The kinetic properties of the enzyme (listed above) suggest that GSH/GSSG ratios in chloroplasts will be very high under both light and dark conditions. This prediction was confirmed experimentally. GSH or GSSG play no part in the light-induced activation of chloroplast fructose diphosphatase or NADP+-glyceraldehyde-3-phosphate dehydrogenase. We suggest that GSH helps to stabilise chloroplast enzymes and may also play a role in removing H2O2. Glucose-6-phosphate dehydrogenase activity may be required in chloroplasts in the dark in order to provide NADPH for glutathione reductase.Abbreviations GSH reduced form of the tripeptide glutathione - GSSG oxidised form of glutathione  相似文献   

5.
The activity of pure calf-liver and Escherichia coli thioredoxin reductases decreased drastically in the presence of NADPH or NADH, while NADP+, NAD+ and oxidized E. coli thioredoxin activated both enzymes significantly, particularly the bacterial one. The loss of activity under reducing conditions was time-dependent, thus suggesting an inactivation process: in the presence of 0.24 mM NADPH the half-lives for the E. coli and calf-liver enzymes were 13.5 and 2 min, respectively. Oxidized E. coli thioredoxin fully protected both enzymes from inactivation, and also promoted their complete reactivation after only 30 min incubation at 30° C. Lower but significant protection and reactivation was also observed with NADP+ and NAD+. EDTA protected thioredoxin reductase from NADPH inactivation to a great degree, thus indicating the participation of metals in the process; EGTA did not protect the enzyme from redox inactivation. Thioredoxin reductase was extensively inactivated by NADPH under aerobic and anaerobic conditions, thus excluding the participation of O2 or oxygen active species in redox inactivation. The loss of thioredoxin reductase activity promoted by NADPH was much faster and complete in the presence of NAD+ glycohydrolase, thus suggesting that inactivation was related to full reduction of the redox-active disulfide. Those results indicate that thioredoxin reductase activity can be modulated in bacteria and mammals by the redox status of NADP(H) and thioredoxin pools, in a similar way to glutathione reductase. This would considerably expand the regulatory potential of the thioredoxin-thioredoxin reductase system with the enzyme being self-regulated by its own substrate, a regulatory protein.Abbreviations DTNB 5,5-dithiobis(2-nitrobenzoate) - EGTA Ethylenglycoltetraacetic Acid - TNB 5-thio-2-nitrobenzoate - Trx Thioredoxin - Trx(SH)2 Reduced Thioredoxin - Trx-S2 Oxidized Thioredoxin  相似文献   

6.
A psychrophilic glutathione reductase from Antarctic ice microalgae Chlamydomonas sp. Strain ICE-L was purified by ammonium sulfate fractionation and three steps of chromatography. The yield was up to 25.1% of total glutathione reductase in the crude enzyme extract. The glutathione reductase activity was characterized by the spectrophotometric method under different conditions. Purified glutathione reductase was separated by SDS-PAGE, which furnished a homogeneous band. The native molecular mass of the enzyme was 115 kDa. Apparent Km values for NADPH and NADH (both at 0.5 mmol L−1 oxidized glutathione) were 22.3 and 83.8 μmol L−1, respectively. It was optimally active at pH 7.5, and it was stable from pH 5 to 9. Its optimum temperature was 25°C, with activity at 0°C 23.5% of the maximum. Its optimum ion strength and optimum Mg2+ were 50–90 and 7.5 mmol L−1, respectively. Ca2+, Mg2+, and cysteine substantially increased the activity of the enzyme but chelating agents, heavy metals (Cd2+, Pb2+, Cu2+, Zn2+, etc.), NADPH, and ADP had significant inhibitory effects. This glutathione reductase can be used to study the adaptation and mechanism of catalysis of psychrophilic enzymes, and it has a high potential as an environmental biochemical indicator under extreme conditions.  相似文献   

7.
Native and recombinant malate dehydrogenase (MDH) was characterized from the hyperthermophilic, facultatively autotrophic archaeon Pyrobaculum islandicum. The enzyme is a homotetramer with a subunit mass of 33 kDa. The activity kinetics of the native and recombinant proteins are the same. The apparent K m values of the recombinant protein for oxaloacetate (OAA) and NADH (at 80°C and pH 8.0) were 15 and 86 μM, respectively, with specific activity as high as 470 U mg−1. Activity decreased more than 90% when NADPH was used. The catalytic efficiency of OAA reduction by P. islandicum MDH using NADH was significantly higher than that reported for any other archaeal MDH. Unlike other archaeal MDHs, specific activity of the P. islandicum MDH back-reaction also decreased more than 90% when malate and NAD+ were used as substrates and was not detected with NADP+. A phylogenetic tree of 31 archaeal MDHs shows that they fall into 5 distinct groups separated largely along taxonomic lines suggesting minimal lateral mdh transfer between Archaea.  相似文献   

8.
Wang X  Fang B  Luo J  Li W  Zhang L 《Biotechnology letters》2007,29(9):1409-1412
The xylose reductase (XR) gene (xyl1) from Candida shehatae was cloned and expressed in Escherichia coli, and purified as a His6-tagged fusion protein. The recombinant XR had Km values for NADH than NADPH of 150 μM and 20 μM, respectively. The optimal reaction was at pH 6.5 and 35°C. The enzyme was specific for d-xylese.  相似文献   

9.
J. Herrera  D.J.D. Nicholas 《BBA》1974,368(1):54-60
Nitrate reductase from Nitrobacter agilis was inactivated by NADH (but not by NADPH) in the absence of nitrate.The inactivation of the enzyme by over-reduction with NADH was overcome by oxidizing the reduced enzyme with nitrate, ferricyanide, NAD+ or NADP+.  相似文献   

10.
A flavoenzyme which showed NADPH-cytochrome c reductase (NADPH-cytochrome c oxidoreductase EC 1.6.2.4) and transhydrogenase (NADPH-NAD+ oxidoreductase, EC 1.6.1.1) activities was purified to an electrophoretically homogeneous state from Nitrobacter winogradskyi. The reductase was a flavoprotein which contained one FAD per molecule but no FMN. The oxidized form of the enzyme showed absorption maxima at 272, 375 and 459 nm with a shoulder at 490 nm, its molecular weight was estimated to be 36,000 by SDS polyacrylamide gel electrophoresis, and the enzyme seemed to exist as a dimer in aqueous solution. The enzyme catalyzed reduction of cytochrome c, DCIP and benzylviologen by NADPH, oxidation of NADPH with menadione and duroquinone, and showed transhydrogenase activity. NADH was less effective than NADPH as the electron donor in the reactions catalyzed by the enzyme. The NADPH-reduction catalyzed by the enzyme of N. winogradskyi cytochrome c-550 and horse cytochrome c was stimulated by spinach ferredoxin. The enzyme reduced NADP+ with reduced spinach ferredoxin and benzylviologen radical.Abbreviations DCIP dichlorophenolindophenol - Tris trishydroxy-methylaminomethane - Mops 3-(N-morpholino) propanesulfonic acid - SDS sodium dodecylsufate  相似文献   

11.
Glutathione reductase was purified from iron-grown Thiobacillus ferrooxidas AP19-3 to an electrophoretically homogeneous state. The enzyme had an apparent molecular weight of 100,000 and was composed of two identical subunits of molecular weight (Mrs, 52,000) as estimated by sodium dodecyl sulfate–polyacrylamide gel electrophoresis. A purified enzyme reduced one mole of the oxidized form of glutathione (GSSG) with one mole of NADPH to produce two moles of the reduced form of glutathione (GSH) and one mole of NADP+. The glutathione reductase was most active at pH 6.5 and 40°C, and had an isoelectric point at 5.1. The Michaelis constants of glutathione reductase for GSSG, NADPH, and NADH were 300, 26, and 125 μM, respectively.  相似文献   

12.
Although inhibition of glutathione reductase (GR) has been demonstrated to cause a decrease in reduced glutathione (GSH) and increase in glutathione disulfide (GSSG), a systematic study of the effects of GR inhibition on thiol redox state and related systems has not been noted. By employing a monkey kidney cell line as the cell model and 2-acetylamino-3-[4-(2-acetylamino-2-carboxy-ethylsulfanylthio carbonylamino)phenylthiocarbamoylsulfanyl]propionic acid (2-AAPA) as a GR inhibitor, an investigation of the effects of GR inhibition on cellular thiol redox state and related systems was conducted. Our study demonstrated that, in addition to a decrease in GSH and increase in GSSG, 2-AAPA increased the ratios of NADH/NAD+ and NADPH/NADP+. Significant protein glutathionylation was observed. However, the inhibition did not affect the formation of reactive oxygen species or expression of antioxidant defense enzyme systems [GR, glutathione peroxidase, catalase, and superoxide dismutase] and enzymes involved in GSH biosynthesis [γ-glutamylcysteine synthetase and glutathione synthetase].  相似文献   

13.
Nicotinamide adenine dinucleotide (NAD+/NADH) along with its phosphorylated form (NADP+/NADPH) are two molecules ubiquitously present in all organisms, and they play key roles as cofactors in fundamental catabolic and anabolic processes, respectively. The oxidation of NADPH to NADP+ initiates a cascade of reactions, where a network of molecules is implicated. The molecules of this cascade form a network with eminent translational potential in redox metabolism. A special point of interest is that spectrophotometric assays have been developed both for NADH/NADPH and the molecules directly regulated by them. Therefore, crucial molecules of the NADPH-dependent redox network can be measured, and the results can be used to assess the bioenergetic and/or oxidative stress status. The main aim of this review is to collectively present the NADPH-related molecules, namely NADPH, NADH, NAD+ kinase, NADPH oxidase, peroxiredoxin, thioredoxin, thioredoxin reductase, and nitric oxide synthase, that can be measured in blood and tissues with the use of a spectrophotometer, which is probably the most simple, inexpensive and widely used tool in biochemistry. We are providing the researchers with reliable and valid spectrophotometric assays for the measurement of the most important biomarkers of the NADPH network in blood and other tissues, thus allowing the opportunity to follow the redox changes in response to a stimulus.  相似文献   

14.
Using isolated chloroplasts or purified thylakoids from photoautotrophically grown cells of the chromophytic alga Pleurochloris meiringensis (Xanthophyceae) we were able to demonstrate a membrane bound NAD(P)H dehydrogenase activity. NAD(P)H oxidation was detectable with menadione, coenzyme Q0, decylplastoquinone and decylubiquinone as acceptors in an in vitro assay. K m-values for both pyridine nucleotides were in the molar range (K m[NADH]=9.8 M, K m[NADPH]=3.2 M calculated according to Lineweaver-Burk). NADH oxidation was optimal at pH 9 while pH dependence of NADPH oxidation showed a main peak at 9.8 and a smaller optimum at pH 7.5–8. NADH oxidation could be completely inhibited with rotenone, an inhibitor of mitochondrial complex I dehydrogenase, while NADPH oxidation revealed the typical inhibition pattern upon addition of oxidized pyridine nucleotides reported for ferredoxin: NADP+ reductase. Partly-denaturing gel electrophoresis followed by NAD(P)H dehydrogenase activity staining showed that NADPH and NADH oxidizing proteins had different electrophoretic mobilities. As revealed by denaturing electrophoresis, the NADH oxidizing enzyme had one main subunit of 22 kDa and two further polypeptides of 29 and 44 kDa, whereas separation of the NADPH depending protein yielded five bands of different molecular weight. Measurement of oxygen consumption due to PS I mediated methylviologen reduction upon complete inhibition of PS II showed that the NAD(P)H dehydrogenase is able to catalyze an input of electrons from NADH to the photosynthetic electron transport chain in case of an oxidized plastoquinone-pool. We suggest ferredoxin: NADP+ reductase to be the main NADPH oxidizing activity while a thylakoidal NAD(P)H: plastoquinone oxidoreductase involved in the chlororespiratory pathway in the dark acts mainly as an NADH oxidizing enzyme.Abbreviations Coenzyme Q0-2,3-dimethoxy-5-methyl-1,4-benzoquinone - FNR ferredoxin: NADP+ reductase - MD menadione - MV methylviologen - NDH NAD(P)H dehydrogenase - PQ plastoquinone - PQ10 decylplastoquinone - SDH succinate dehydrogenase - UQ10 decylubiquinone (2,3-dimethoxy-5-methyl-6-decyl-1,4-benzoquinone)  相似文献   

15.
Given the rapid rise in antibiotic resistance, including methicillin resistance in Staphylococcus aureus (MRSA), there is an urgent need to characterize novel drug targets. Enzymes of the lysine biosynthesis pathway in bacteria are examples of such targets, including dihydrodipicolinate reductase (DHDPR, E.C. 1.3.1.26), which is the product of an essential bacterial gene. DHDPR catalyzes the NAD(P)H-dependent reduction of dihydrodipicolinate (DHDP) to tetrahydrodipicolinate (THDP) in the lysine biosynthesis pathway. We show that MRSA–DHDPR exhibits a unique nucleotide specificity utilizing NADPH (Km = 12 μM) as a cofactor more effectively than NADH (Km = 26 μM). However, the enzyme is inhibited by high concentrations of DHDP when using NADPH as a cofactor, but not with NADH. Isothermal titration calorimetry (ITC) studies reveal that MRSA–DHDPR has ∼20-fold greater binding affinity for NADPH (Kd = 1.5 μM) relative to NADH (Kd = 29 μM). Kinetic investigations in tandem with ITC studies show that the enzyme follows a compulsory-order ternary complex mechanism; with inhibition by DHDP through the formation of a nonproductive ternary complex with NADP+. This work describes, for the first time, the catalytic mechanism and cofactor preference of MRSA–DHDPR, and provides insight into rational approaches to inhibiting this valid antimicrobial target.  相似文献   

16.
Glucose-6-phosphate dehydrogenase (d-glucose-6-phosphate: NADP+ l-oxidoreductase EC 1.1.1.49) isolated from Paracoccus denitrificans grown on glucose/nitrate exhibits both NAD+-and NADP+-linked activities. Both activities have a pH optimum of pH 9.6 (Glycine/NaOH buffer) and neither demonstrates a Mg2+ requirement. Kinetics for both NAD(P)+ and glucose-6-phosphate were investigated. Phosphoenolpyruvate inhibits both activities in a competitive manner with respect to glucose-6-phosphate. ATP inhibits the NAD+-linked activity competitively with respect to glucose-6-phosphate but has no effect on the NADP+-linked activity. Neither of the two activities are inhibited by 100 M NADH but both are inhibited by NADPH. The NAD+-linked activity is far more sensitive to inhibition by NADPH than the NADP+-linked activity.  相似文献   

17.
NADP is a key electron carrier for a broad spectrum of redox reactions, including photosynthesis. Hence, chloroplastic NADP status, as represented by redox status (ratio of NADPH to NADP+) and pool size (sum of NADPH and NADP+), is critical for homeostasis in photosynthetic cells. However, the mechanisms and molecules that regulate NADP status in chloroplasts remain largely unknown. We have now characterized an Arabidopsis mutant with imbalanced NADP status (inap1), which exhibits a high NADPH/NADP+ ratio and large NADP pool size. inap1 is a point mutation in At2g04700, which encodes the catalytic subunit of ferredoxin/thioredoxin reductase. Upon illumination, inap1 demonstrated earlier increases in NADP pool size than the wild type did. The mutated enzyme was also found in vitro to inefficiently reduce m‐type thioredoxin, which activates Calvin cycle enzymes, and NADP‐dependent malate dehydrogenase to export reducing power to the cytosol. Accordingly, Calvin cycle metabolites and amino acids diminished in inap1 plants. In addition, inap1 plants barely activate NADP‐malate dehydrogenase, and have an altered redox balance between the chloroplast and cytosol, resulting in inefficient nitrate reduction. Finally, mutants deficient in m‐type thioredoxin exhibited similar light‐dependent NADP dynamics as inap1. Collectively, the data suggest that defects in ferredoxin/thioredoxin reductase and m‐type thioredoxin decrease the consumption of NADPH, leading to a high NADPH/NADP+ ratio and large NADP pool size. The data also suggest that the fate of NADPH is an important influence on NADP pool size.  相似文献   

18.
Nitrate reductase (NR) (EC 1.6.6.2) from Chlorella variegata 211/10d has been purified by blue sepharose affinity chromatography. The enzyme can utilise NADH or NADPH for nitrate reduction with apparent K m values of 11.5 M and 14.5 M, respectively. Apparent K m values for nitrate are 0.13 mM (NADH-NR) and 0.14 mM (NADPH-NR). The diaphorase activity of the enzyme is inhibited strongly by parachloromercuribenzoic acid; NADH or NADPH protects the enzyme against this inhibition. NR proper activity of the enzyme is partially inactive after extraction and may be activated after the addition of ferricyanide. The addition of NAD(P)H and cyanide causes a reversible inactivation of the NR proper activity although preincubation with either NADH or NADH and ADP has no significant effect.Abbreviations NR Nitrate reductase - FAD Flavin-adenine dinucleotide - FMN Riboflavin 5-phosphate - p-CMB para-Chloromercuribenzoic - BV Benzyl viologen  相似文献   

19.
Igamberdiev AU  Bykova NV  Hill RD 《Planta》2006,223(5):1033-1040
NADH-dependent NO scavenging in barley extracts is linked to hemoglobin (Hb) expression and is inhibited by SH-reagents. Barley Hb has a single cysteine residue. To determine whether this cysteine was critical for NO scavenging, barley Hb and a mutated version, in which the single Cys79 was replaced by Ser, were over-expressed in Escherichia coli and purified to near homogeneity. The purified proteins exhibited very low NO-scavenging activity (12–14 nmol min−1 mg−1 protein) in the presence of NADH or NADPH. This activity was insensitive to SH-reagents. Addition of an extract from barley roots to either of the purified proteins resulted in high NADH-dependent NO turnover in a reaction that was sensitive to SH-reagents. A protein was purified from barley roots and identified by mass-spectrometry analysis as a cytosolic monodehydroascorbate reductase. It efficiently supported NADH-dependent NO scavenging in the presence of either native or mutated barley Hb. Ascorbate strongly facilitated the rate of metHb reduction. The K m for Hb was 0.3 μM, for ascorbate 0.6 mM and for NADH 4 μM. The reaction in the presence of monodehydroascorbate reductase was sensitive to SH-reagents with either form of the Hb. We conclude that metHb reduction and NO turnover do not involve direct participation of the Cys79 residue of barley Hb. NO scavenging is facilitated by monodehydroascorbate reductase mediating a coupled reaction involving ferric Hb reduction in the presence of ascorbate and NADH.  相似文献   

20.
d-Glucose dehydrogenase [β-d-glucosc: NAD(P) oxidoreductase (EC 1.1.1.47)] was synthesized derepressively in a mutant of a Bacillus species which was isolated as an improved strain for d-ribose production. The enzyme was very unstable and inactivated during storage or column chromatography. The inactivation was prevented in the presence of NAD+, NADP+ or certain salts. The inactive enzyme was reactivated by the addition of NAD+, NADH, NADP+, NADPH, AMP, ADP, ATP or certain salts. The molecular weights of the inactive and active form of the enzyme were estimated to be about 45,000 and 80,000, respectively, by Sephadex G–150 gel filtration. Thus, it seems that the enzyme activity is regulated by monomer-dimer interconversion of the enzyme molecule.  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