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 15 毫秒
1.
Numerous physiological aldehydes besides glucose are substrates of aldose reductase, the first enzyme of the polyol pathway which has been implicated in the etiology of diabetic complications. The 2-oxoaldehyde methylglyoxal is a preferred substrate of aldose reductase but is also the main physiological substrate of the glutathione-dependent glyoxalase system. Aldose reductase catalyzes the reduction of methylglyoxal efficiently (kcat=142 min−1 and kcat/Km=1.8×107 M−1 min−1). In the presence of physiological concentrations of glutathione, methylglyoxal is significantly converted into the hemithioacetal, which is the actual substrate of glyoxalase-I. However, in the presence of glutathione, the efficiency of reduction of methylglyoxal, catalyzed by aldose reductase, also increases. In addition, the site of reduction switches from the aldehyde to the ketone carbonyl. Thus, glutathione converts aldose reductase from an aldehyde reductase to a ketone reductase with methylglyoxal as substrate. The relative importance of aldose reductase and glyoxalase-I in the metabolic disposal of methylglyoxal is highly dependent upon the concentration of glutathione, owing to the non-catalytic pre-enzymatic reaction between methylglyoxal and glutathione.  相似文献   

2.
Abstract— The presence of a nonspecific NADH-linked aldehyde reductase was demonstrated in various regions of bovine brain in vitro. With m-nitrobenzaldehyde as substrate, the rate of NADH oxidation was approximately 4 nmol.min-1.(mg of protein)-1 in the cerebellum, pons and medulla; but somewhat lower rates [2–3 nmol.min-1.(mg of protein)-l] were obtained in the other areas of the brain examined. The enzyme was localized primarily in the soluble, supernatant fraction of rat brain homogenates. The enzyme from the supernatant fluid fraction of bovine brain was purified approximately 350-fold by ammonium sulphate fractionation and chromatography on calcium phosphate-gel, DEAE-cellulose and Sephadex G200 columns. The partially purified enzyme catalysed the reduction of a number of aldehydes, including substituted benzaldehydes and aliphatic aldehydes of intermediate chain lengths. Short chain aliphatic aldehydes, such as acetaldehyde, were not reduced by the enzyme and butyraldehyde was a poor substrate. With m-nitrobenzaldehyde as substrate, NADH was oxidized at an approximately 10-fold faster rate than NADPH. The pH optimum for the enzyme was 6.75 for aldehyde reduction, whereas the rate of oxidation of m-nitrobenzylalcohol was optimal at pH 10.0 with NAD as the co-substrate. Km and K3 values ranged from 10 μM to 10 mM for various aldehydes and from 10 to 30 μM for the cofactors. Oxidation of NADH by the partially purified enzyme was not inhibited by 10m pyrazole or by 1 mM phenobarbital. However, the enzyme activity was inhibited by approximately 60 percent by 1 mM chlorpromazine or by 5 mM 1,10-orthophenanthroline. Our data demonstrate that the enzyme is not only separable from the NADPH-linked aldehyde reductase described previously by TABAKOFF and ERWIN, but also is quite different in substrate specificity and inhibitor sensitivity from the ‘classical’, pyrazole-sensitive, NAD- linked alcohol dehydrogenase (EC 1.1.1.1).  相似文献   

3.
A novel whole-cell biocatalyst with high allylic alcohol-oxidizing activities was screened and identified as Yokenella sp. WZY002, which chemoselectively reduced the C=O bond of allylic aldehydes/ketones to the corresponding α,β-unsaturated alcohols at 30°C and pH 8.0. The strain also had the capacity of stereoselectively reducing aromatic ketones to (S)-enantioselective alcohols. The enzyme responsible for the predominant allylic/benzyl alcohol dehydrogenase activity was purified to homogeneity and designated YsADH (alcohol dehydrogenase from Yokenella sp.), which had a calculated subunit molecular mass of 36,411 Da. The gene encoding YsADH was subsequently expressed in Escherichia coli, and the purified recombinant YsADH protein was characterized. The enzyme strictly required NADP(H) as a coenzyme and was putatively zinc dependent. The optimal pH and temperature for crotonaldehyde reduction were pH 6.5 and 65°C, whereas those for crotyl alcohol oxidation were pH 8.0 and 55°C. The enzyme showed moderate thermostability, with a half-life of 6.2 h at 55°C. It was robust in the presence of organic solvents and retained 87.5% of the initial activity after 24 h of incubation with 20% (vol/vol) dimethyl sulfoxide. The enzyme preferentially catalyzed allylic/benzyl aldehydes as the substrate in the reduction of aldehydes/ketones and yielded the highest activity of 427 U mg−1 for benzaldehyde reduction, while the alcohol oxidation reaction demonstrated the maximum activity of 79.9 U mg−1 using crotyl alcohol as the substrate. Moreover, kinetic parameters of the enzyme showed lower Km values and higher catalytic efficiency for crotonaldehyde/benzaldehyde and NADPH than for crotyl alcohol/benzyl alcohol and NADP+, suggesting the nature of being an aldehyde reductase.  相似文献   

4.
A putative aldehyde reductase gene from Oceanospirillum sp. MED92 was overexpressed in Escherichia coli. The recombinant protein (OsAR) was characterized as a monomeric NADPH-dependent aldehyde reductase. The kinetic parameters Km and kcat of OsAR were 0.89 ± 0.08 mM and 11.07 ± 0.99 s−1 for benzaldehyde, 0.04 ± 0.01 mM and 6.05 ± 1.56 s−1 for NADPH, respectively. This enzyme exhibited high activity toward a variety of aromatic and aliphatic aldehydes, but no activity toward ketones. As such, it catalyzed the chemoselective reduction of aldehydes in the presence of ketones, as demonstrated by the reduction of 4-acetylbenzaldehyde or the mixture of hexanal and 2-nonanone, showing the application potential of this marine enzyme in such selective reduction of synthetic importance.  相似文献   

5.
The 2.2 Å X-ray crystal structure of Candida tenuis xylose reductase (AKR2B5) bound with NADP+ reveals that Phe-114 contributes to the substrate binding pocket of the enzyme. In the related human aldose reductase (AKR1B1), this phenylalanine is replaced by a tryptophan. The side chain of Trp was previously implicated in forming a hydrogen bond with bound substrate or inhibitor. The apparent Michaelis constant of AKR2B5 for xylose (Km≈90 mM) is 60 times that of AKR1B1, perhaps because critical enzyme–substrate interactions of Trp are not available to Phe-114. We, therefore, prepared a Phe-114→Trp mutant (F114W) of AKR2B5, to mimic the aldose reductase relationship in xylose reductase. Detailed analysis of the kinetic consequences in purified F114W revealed that the Km values for xylose and xylitol at pH 7.0 and 25°C were increased 5.1- and 4.4-fold, respectively, in the mutant compared with the wild-type. Turnover numbers (kcat) of F114W for xylose reduction and xylitol oxidation were half those of the wild-type. Apparent dissociation constants of NADH (KiNADH=44 µM) and NAD+ (KiNAD+=177 µM) were increased 1.6- and 1.4-fold in comparison with values of KiNADH and KiNAD+ for the wild-type, respectively. Catalytic efficiencies (kcat/Km) for NADH-dependent reduction of different aldehydes were between 3.1- and 31.5-fold lower than the corresponding kcat/Km values of the wild-type. Therefore, replacement of Phe-114 with Trp weakens rather than strengthens apparent substrate binding by AKR2B5, suggesting that xylose reductase exploits residue 114 in a different manner from aldose reductase.  相似文献   

6.
Erythritol biosynthesis is catalyzed by erythrose reductase, which converts erythrose to erythritol. Erythrose reductase, however, has never been characterized in terms of amino acid sequence and kinetics. In this study, NAD(P)H-dependent erythrose reductase was purified to homogeneity from Candida magnoliae KFCC 11023 by ion exchange, gel filtration, affinity chromatography, and preparative electrophoresis. The molecular weights of erythrose reductase determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel filtration chromatography were 38,800 and 79,000, respectively, suggesting that the enzyme is homodimeric. Partial amino acid sequence analysis indicates that the enzyme is closely related to other yeast aldose reductases. C. magnoliae erythrose reductase catalyzes the reduction of various aldehydes. Among aldoses, erythrose was the preferred substrate (Km = 7.9 mM; kcat/Km = 0.73 mM−1 s−1). This enzyme had a dual coenzyme specificity with greater catalytic efficiency with NADH (kcat/Km = 450 mM−1 s−1) than with NADPH (kcat/Km = 5.5 mM−1 s−1), unlike previously characterized aldose reductases, and is specific for transferring the 4-pro-R hydrogen of NADH, which is typical of members of the aldo/keto reductase superfamily. Initial velocity and product inhibition studies are consistent with the hypothesis that the reduction proceeds via a sequential ordered mechanism. The enzyme required sulfhydryl compounds for optimal activity and was strongly inhibited by Cu2+ and quercetin, a strong aldose reductase inhibitor, but was not inhibited by aldehyde reductase inhibitors and did not catalyze the reduction of the substrates for carbonyl reductase. These data indicate that the C. magnoliae erythrose reductase is an NAD(P)H-dependent homodimeric aldose reductase with an unusual dual coenzyme specificity.  相似文献   

7.
The gene xylBADP1 from Acinetobacter baylyi ADP1 (gene annotation number ACIAD1578), coding for a putative aryl alcohol dehydrogenase, was heterologously expressed in Escherichia coli BL21(DE3). The respective aryl alcohol dehydrogenase was purified by fast protein liquid chromatography to apparent electrophoretic homogeneity. The predicted molecular weight of 39,500 per subunit was confirmed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. According to the native Mw as determined by gel filtration, the enzyme forms dimers and therefore seems to be XylB related. The enzyme showed the highest activity at 40°C. For both the reduction and the oxidation reactions, the pH for optimum activity was 6.5. The enzyme was NADH dependent and able to reduce medium- to long-chain n-alkylaldehydes, methyl-branched aldehydes, and aromatic aldehydes, with benzaldehyde yielding the highest activity. The oxidation reaction with the corresponding alcohols showed only 2.2% of the reduction activity, with coniferyl alcohol yielding the highest activity. Maximum activities for the reduction and the oxidation reaction were 104.5 and 2.3 U mg−1 of protein, respectively. The enzyme activity was affected by low concentrations of Ag+ and Hg2+ and high concentrations of Cu2+, Zn2+, and Fe2+. The gene xylBADP1 seems to be expressed constitutively and an involvement in coniferyl alcohol degradation is suggested. However, the enzyme is most probably not involved in the degradation of benzyl alcohol, anisalcohol, salicyl alcohol, vanillyl alcohol, cinnamyl alcohol, or aliphatic and isoprenoid alcohols.  相似文献   

8.
Synthesis of long-chain fatty alcohols in preputial glands of mice is catalyzed by an NADPH-dependent acyl coenzyme A (CoA) reductase located in microsomal membranes; sensitivity to trypsin digestion indicates that the reductase is on the cytoplasmic side of the membrane. Results with pyrazole and phenobarbital demonstrate the reaction is not catalyzed by a nonspecific alcohol dehydrogenase or an aldehyde reductase. Acyl-CoA reductase activity is sensitive to sulfhydryl and serine reagent modification, is stimulated by bovine serum albumin, and produces an aldehyde intermediate. The activity is extremely detergent sensitive and cannot be restored even after removal of the detergents. Phospholipase C or asolectin treatment does not release the acyl-CoA reductase from microsomal membranes, but causes a significant decrease in the activity recovered in the membrane pellet. Glycerol does not solubilize the reductase activity, nor does 3.0 m NaCl; however, the combination of glycerol and 3.0 m NaCl did release about 50% of the acyl-CoA reductase from the microsomal pellet. Substrate concentration curves obtained in the presence or absence of bovine serum albumin show significant differences in enzyme activities. The reductase is sensitive to the concentration of palmitoyl-CoA and is progressively inhibited at levels beyond the critical micellar concentration of the substrate. The apparent Km for acyl-CoA reductase is 14 μm; however, the maximum velocity varies with the concentration of albumin used. Expression of enzyme activity in delipidated microsomes requires specific phospholipids, which suggests that in vivo regulation of acyl-CoA reductase activity could be achieved through modifications in membrane lipid composition.  相似文献   

9.
The recently identified benzoate oxidation (box) pathway in Burkholderia xenovorans LB400 (LB400 hereinafter) assimilates benzoate through a unique mechanism where each intermediate is processed as a coenzyme A (CoA) thioester. A key step in this process is the conversion of 3,4-dehydroadipyl-CoA semialdehyde into its corresponding CoA acid by a novel aldehyde dehydrogenase (ALDH) (EC 1.2.1.x). The goal of this study is to characterize the biochemical and structural properties of the chromosomally encoded form of this new class of ALDHs from LB400 (ALDHC) in order to better understand its role in benzoate degradation. To this end, we carried out kinetic studies with six structurally diverse aldehydes and nicotinamide adenine dinucleotide (phosphate) (NAD + and NADP +). Our data definitively show that ALDHC is more active in the presence of NADP + and selective for linear medium-chain to long-chain aldehydes. To elucidate the structural basis for these biochemical observations, we solved the 1.6-Å crystal structure of ALDHC in complex with NADPH bound in the cofactor-binding pocket and an ordered fragment of a polyethylene glycol molecule bound in the substrate tunnel. These data show that cofactor selectivity is governed by a complex network of hydrogen bonds between the oxygen atoms of the 2′-phosphoryl moiety of NADP + and a threonine/lysine pair on ALDHC. The catalytic preference of ALDHC for linear longer-chain substrates is mediated by a deep narrow configuration of the substrate tunnel. Comparative analysis reveals that reorientation of an extended loop (Asn478-Pro490) in ALDHC induces the constricted structure of the substrate tunnel, with the side chain of Asn478 imposing steric restrictions on branched-chain and aromatic aldehydes. Furthermore, a key glycine (Gly104) positioned at the mouth of the tunnel allows for maximum tunnel depth required to bind medium-chain to long-chain aldehydes. This study provides the first integrated biochemical and structural characterization of a box-pathway-encoded ALDH from any organism and offers insight into the catalytic role of ALDHC in benzoate degradation.  相似文献   

10.
Long chain fatty alcohols have wide application in chemical industries and transportation sector. There is no direct natural reservoir for long chain fatty alcohol production, thus many groups explored metabolic engineering approaches for its microbial production. Escherichia coli has been the major microbial platform for this effort, however, terminal endogenous enzyme responsible for converting fatty aldehydes of chain length C14-C18 to corresponding fatty alcohols is still been elusive. Through our in silico analysis we selected 35 endogenous enzymes of E. coli having potential of converting long chain fatty aldehydes to fatty alcohols and studied their role under in vivo condition. We found that deletion of ybbO gene, which encodes NADP+ dependent aldehyde reductase, led to >90% reduction in long chain fatty alcohol production. This feature was found to be strain transcending and reinstalling ybbO gene via plasmid retained the ability of mutant to produce long chain fatty alcohols. Enzyme kinetic study revealed that YbbO has wide substrate specificity ranging from C6 to C18 aldehyde, with maximum affinity and efficiency for C18 and C16 chain length aldehyde, respectively. Along with endogenous production of fatty aldehyde via optimized heterologous expression of cyanobaterial acyl-ACP reductase (AAR), YbbO overexpression resulted in 169 mg/L of long chain fatty alcohols. Further engineering involving modulation of fatty acid as well as of phospholipid biosynthesis pathway improved fatty alcohol production by 60%. Finally, the engineered strain produced 1989 mg/L of long chain fatty alcohol in bioreactor under fed-batch cultivation condition. Our study shows for the first time a predominant role of a single enzyme in production of long chain fatty alcohols from fatty aldehydes as well as of modulation of phospholipid pathway in increasing the fatty alcohol production.  相似文献   

11.
3-Hydroxypropionic acid (3-HP), a versatile and valuable platform chemical, has diverse industrial applications; but its biological production from glycerol is often limited by the capability of the enzyme aldehyde dehydrogenase (ALDH) to convert an intermediary compound, 3-hydroxypropionaldehyde (3-HPA), to 3-HP. In this study, we report a new ALDH, PuuC, from Klebsiella pneumoniae DSM 2026, that efficiently converts 3-HPA to 3-HP. The identified gene puuC was cloned, expressed in Escherichia coli, purified, and characterized for its properties. The recombinant enzyme with a molecular weight of 53.8 kDa exhibited broad substrate specificity for various aliphatic aldehydes, especially C2–C5 aldehydes. NAD+ was the preferred coenzyme for the oxidation of most aliphatic and aromatic aldehydes tested. The optimum pH and temperature for PuuC activity were pH 8.0 and 45°C. The K m values for 3-HPA and NAD+ were 0.48 and 0.09 mM, respectively. The activity of PuuC was enhanced in the presence of reducing agents such as 2-mercaptoethanol or dithiothreitol, while several metal ions, particularly Hg2+, Ag+, and Cu2+ inhibited its activity. The predicted structure of PuuC indicated the presence of K191 and E194 in close proximity to the glycine motif, suggesting that PuuC belongs to class 2 ALDHs.  相似文献   

12.
Bromphenol blue, which was reduced with dithionite, was found to support nitrate reduction catalyzed by squash NADH:nitrate reductase at a rate about 5 times greater than NADH with freshly prepared enzyme and 10 times or more with enzyme having been frozen and thawed. Kinetic analysis of bromphenol blue as a substrate for squash nitrate reductase yielded apparent Km values of 60 micromolar for bromphenol blue at 10 millimolar nitrate and 500 micromolar for nitrate at 0.2 millimolar bromphenol blue. With the same preparation of enzyme the apparent Km values were 9 micromolar for NADH at 10 millimolar nitrate and 50 micromolar nitrate at 0.1 millimolar NADH. Bromphenol blue was found to be a noncompetitive inhibitor versus NADH with a Ki of 0.3 millimolar. When squash NADH:nitrate reductase activity was inactivated with p-hydroxymercuribenzoate or denatured by heating at 40°C, the bromphenol blue nitrate reductase activity was not lost. These results were taken to indicate that bromphenol blue and NADH donated electrons to nitrate reductase at different sites. When monoclonal antibodies prepared against corn and squash nitrate reductases were used to inhibit the nitrate reductase activities supported by NADH, bromphenol blue, and methyl viologen, differential inhibition was found which tended to indicate that the three electron donors were interacting with the enzyme at different sites. One monoclonal antibody prepared against squash nitrate reductase inhibited all three activities of both corn and squash nitrate reductase. It appears this antibody may bind to a highly conserved antigenic site in the nitrate binding region of the enzyme.  相似文献   

13.
An enzyme assay was developed to determine the activities of methyl chloride dehalogenase and O-demethylase of the homoacetogen strain MC. The formation of methyl tetrahydrofolate from tetrahydrofolate and methyl chloride or from tetrahydrofolate and vanillate was coupled to the oxidation of methyl tetrahydrofolate to methylene tetrahydrofolate mediated by methylene tetrahydrofolate reductase purified from Peptostreptococcus productus (strain Marburg) and to the subsequent oxidation of methylene tetrahydrofolate to methenyl tetrahydrofolate catalyzed by methylene tetrahydrofolate dehydrogenase purified from the same organism. To drive the endergonic methyl tetrahydrofolate oxidation with NAD+ as an electron acceptor, the NADH formed in this reaction was reoxidized in the exergonic lactate dehydrogenase reaction. The formation of NADPH and methenyl tetrahydrofolate in the methylene tetrahydrofolate dehydrogenase reaction was followed photometrically at 350 nm; ε350 was about 29.5 mM–1cm–1 (pH 6.5). Using the coupled enzyme assay, the cofactor requirements, the apparent kinetic parameters, the pH and temperature optima of both enzymes, and the effect of inhibitors were determined. The activity of methyl chloride dehalogenase and of O-demethylase was dependent on the presence of ATP; arsenate severely inhibited both enzyme activities in the absence of ATP. The coupled enzyme assay described allows purification and characterization of methyl chloride dehalogenase and O-demethylase and is also appropriate for the enzymatic determination of methyl tetrahydrofolate. Received: 2 August 1995 / Accepted: 28 September 1995  相似文献   

14.
Procedures for the purification of an aldehyde dehydrogenase from extracts of the obligate methylotroph, Methylomonas methylovora are described. The purified enzyme is homogeneous as judged from polyacrylamide gel electrophoresis. In the presence of an artificial electron acceptor (phenazine methosulfate), the purified enzyme catalyzes the oxidation of straight chain aldehydes (C1-C10 tested), aromatic aldehydes (benzaldehyde, salicylaldehyde), glyoxylate, and glyceraldehyde. Biological electron acceptors such as NAD+, NADP+, FAD, FMN, pyridoxal phosphate, and cytochrome c cannot act as electron carriers. The activity of the enzyme is inhibited by sulfhydryl agents [p-chloromercuribenzoate, N-ethylmaleimide and 5,5-dithiobis (2-nitrobenzoic acid)], cuprous chloride, and ferrour nitrate. The molecular weight of the enzyme as estimated by gel filtration is approximately 45000 and the subunit size determined by sodium dodecyl sulfate-gel electrophoresis is approximately 23000. The purified enzyme is light brown and has an absorption peak at 410 nm. Reduction of enzyme with sodium dithionite or aldehyde substrate resulted in the appearance of peaks at 523 nm and 552 nm. These results suggest that the enzyme is a hemoprotein. There was no evidence that flavins were present as prosthetic group. The amino acid composition of the enzyme is also presented.Non-Standard Abbreviations PMS phenazine methosulfate - DCPIP 2,6-dichlorophenol indophenol - DEAE diethylaminoethyl  相似文献   

15.
Alcohol dehydrogenases are highly diverse enzymes catalysing the interconversion of alcohols and aldehydes or ketones. Due to their versatile specificities, these biocatalysts are of great interest for industrial applications. The adh3-gene encoding a group III alcohol dehydrogenase was isolated from the gram-positive bacterium Oenococcus oeni and was characterised after expression in the heterologous host Escherichia coli. Adh3 has been identified by genome BLASTP analyses using the amino acid sequence of 1,3-propanediol dehydrogenase DhaT from Klebsiella pneumoniae and group III alcohol dehydrogenases with known activity towards 1,3-propanediol as target sequences. The recombinant protein was purified in a two-step column chromatography approach. Crystal structure determination and biochemical characterisation confirmed that Adh3 forms a Ni2+-containing homodimer in its active form. Adh3 catalyses the interconversion of ethanol and its corresponding aldehyde acetaldyhyde and is also capable of using other alcoholic compounds as substrates, such as 1,3-propanediol, 1,2-propanediol and 1-propanol. In the presence of Ni2+, activity increases towards 1,3-propanediol and 1,2-propanediol. Adh3 is strictly dependent on NAD+/NADH, whereas no activity has been observed with NADP+/NADPH as co-factor. The enzyme exhibits a specific activity of 1.1 U/mg using EtOH as substrate with an optimal pH value of 9.0 for ethanol oxidation and 8.0 for aldehyde reduction. Moreover, Adh3 exhibits tolerance to several metal ions and organic solvents, but is completely inhibited in the presence of Zn2+. The present study demonstrates that O. oeni is a group III alcohol dehydrogenase with versatile substrate specificity, including Ni2+-dependent activity towards 1,3-propanediol.  相似文献   

16.
It has been shown recently that in spinach leaves (Spinacia oleracea) net photosynthesis and nitrate reduction are closely linked: when net photosynthesis was low because of stomatal closure, rates of nitrate reduction decreased (WM Kaiser, J Förster [1989] Plant Physiol 91: 970-974). Here we present evidence that photosynthesis regulates nitrate reduction by modulating nitrate reductase activity (NRA, EC 1.6.6.1). When spinach leaves were exposed to low CO2 in the light, extractable NRA declined rapidly with a half-time of 15 minutes. The inhibition was rapidly reversed when leaves were brought back to air. NRA was also inhibited when leaves were wilted in air; this inhibition was due to decreased CO2 supply as a consequence of stomatal closure. The modulation of NRA was stable in vitro. It was not reversed by gel filtration. In contrast, the in vitro inhibition of nitrate reductase (NR) by classical inhibitors such as cyanide, hydroxylamin, or NADH disappeared after removal of free inhibitors by gel filtration. The negative modulation of NRA in —CO2-treated leaves became manifest as a decrease in total enzyme activity only in the presence of free Mg2+ or Ca2+. Mg2+ concentrations required for observing half-maximal inhibition were about 1 millimolar. In the presence of EDTA, the enzyme activity was always high and rather independent of the activation status of the enzyme. NRA was also independent of the pH in the range from pH 7 to pH 8, at saturating substrate and Mg2+ concentrations. The apparent substrate affinities of NR were hardly affected by the in vivo modulation of NR. Only Vmax changed.  相似文献   

17.
Production of the compatible solute glycine betaine from its precursors choline or glycine betaine aldehyde confers a considerable level of tolerance against high osmolarity stress to the soil bacterium Bacillus subtilis. The glycine betaine aldehyde dehydrogenase GbsA is an integral part of the osmoregulatory glycine betaine synthesis pathway. We strongly overproduced this enzyme in an Escherichia coli strain that expressed a plasmid-encoded gbsA gene under T7φ10 control. The recombinant GbsA protein was purified 23-fold to apparent homogeneity by fractionated ammonium sulfate precipitation, ion-exchange chromatography on Q-Sepharose, and subsequent hydrophobic interaction chromatography on phenyl-Sepharose. Molecular sieving through Superose 12 and sedimentation centrifugation through a glycerol gradient suggested that the native enzyme is a homodimer with 53.7-kDa subunits. The enzyme was specific for glycine betaine aldehyde and could use both NAD+ and NADP+ as cofactors, but NAD+ was strongly preferred. A kinetic analysis of the GbsA-mediated oxidation of glycine betaine aldehyde to glycine betaine revealed K m values of 125 μM and 143 μM for its substrates glycine betaine aldehyde and NAD+, respectively. Low concentrations of salts stimulated the GbsA activity, and the enzyme was highly tolerant of high ionic conditions. Even in the presence of 2.4 M KCl, 88% of the initial enzymatic activity was maintained. B. subtilis synthesizes high levels of proline when grown at high osmolarity, and the presence of this amino acid strongly stimulated the GbsA activity in vitro. The enzyme was stimulated by moderate concentrations of glycine betaine, and its activity was highly tolerant against molar concentrations of this osmolyte. The high salt tolerance and its resistance to its own reaction product are essential features of the GbsA enzyme and ensure that B. subtilis can produce high levels of the compatible solute glycine betaine under conditions of high osmolarity stress. Received: 2 May 1997 / Accepted: 2 July 1997  相似文献   

18.
The enzymatic complex nitrate reductase from Spinacea oleracea is inactivated by NADH or NADPH and by simple thiols. The inactivation affects FNH2-nitrate reductase but not NADH-diaphorase. Reactivation can be achieved by addition of ferricyanide. The extent of inactivation by dithioerythritol is increased by NAD+, but not by NADP+. Nitrate protects against inactivation by NADH or NADPH, and abolishes the effect of NAD+ on the inactivation by dithioerythritol. The NAD(P)H-inactivation of nitrate reductase requires that the diaphorase moiety of the complex be functional. However, there is no proportionality between NADH-diaphorase or NADH-nitrate reductase activities and the susceptibility of the enzymatic preparation to NADH or NADPH. It seems likely that the nitrate reductase complex contains a specific regulatory site, different from the catalytic site, the reduction of which is accompanied by the production of an inactive form of the complex.  相似文献   

19.
Oxidative deamination of putrescine, the precursor of polyamines, gives rise to γ-aminobutyraldehyde (ABAL). In this study an aldehyde dehydrogenase, active on ABAL, has been purified to electrophoretic homogeneity from rat liver cytoplasm and its kinetic behaviour investigated. The enzyme is a dimer with a subunit molecular weight of 51,000. It is NAD+-dependent, active only in the presence of sulphhydryl compounds and has a pH optimum in the range 7.3–8.4. Temperatures higher than 28°C promote slow activation and the process is favoured by the presence of at least one substrate. Km for aliphatic aldehydes decreases from 110 μM for ABAL and acetaldehyde to 2–3 μM for capronaldehyde. The highest relative V-values have been observed with ABAL (100) and isobutyraldehyde (64), and the lowest with acetaldehyde (14). Affinity for NAD+ is affected by the aldehyde present at the active site: Km for NAD+ is 70 μM with ABAL, 200 μM with isobutyraldehyde and capronaldehyde, and>800 μM with acetaldehyde. The kinetic behaviour at 37°C is quite complex; according to enzymatic models, NAD+ activates the enzyme (Kact 500 μM) while NADH competes for the regulatory site (Kin 70 μM). In the presence of high NAD+ concentrations (4 mM), ABAL promotes further activation by binding to a low-affinity regulatory site (Kact 10 mM). The data show that the enzyme is probably an E3 aldehyde dehydrogenase, and suggest that it can effectively metabolize aldehydes arising from biogenic amines.  相似文献   

20.
A novel putative aldehyde dehydrogenase (ALDH) gene aldh1413 from Flammeovirga pacifica isolated from deep sea sediment was cloned, expressed, and characterized. The molecular weight of the ALDH1413 (479 amino acids) was estimated by SDS-PAGE to be 53 kDa. The optimum temperature and pH for ALDH1413 were 35°C and 9.0, respectively. In the presence of either NAD+ or NADP+, the enzyme could oxidize a number of aliphatic aldehydes, particularly C3-and C5-aliphatic aldehydes and aromatic aldehydes such as benzaldehyde, which indicates that the enzyme belongs to broad-specific (ALDH) superfamily. Steady-state kinetic study revealed that ALDH1413 had a K M value of 0.545 mM and a k cat value of 7.48 s?1 when propionaldehyde was used as the substrate. The Na+ could enhance ALDH1413 activity, which indicated it might be adapt to its habitat, marine environment.  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