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.
Mammalian dimeric dihydrodiol dehydrogenase is identical with d-xylose dehydrogenase and belongs to a protein family with prokaryotic proteins including glucose-fructose oxidoreductase. Of the conserved residues in this family, either His-79 or Tyr-180 of d-xylose/dihydrodiol dehydrogenase has been proposed to be involved in the catalytic function. Site-directed mutagenesis was used to examine the roles of the two residues of the monkey enzyme. A mutant, Y180F, was almost inactive, but, similarly to the wild-type enzyme, exhibited high affinity for NADP(H) and fluorescence energy transfer upon binding of NADPH. The H79Q mutation had kinetically largest effects on K(d) (>7-fold increase) and K(m) (>25-fold increase) for NADP(H), and eliminated the fluorescence energy transfer. Interestingly, the dehydrogenase activity of this mutant was potently inhibited with a 190-fold increase in the K(m) for NADP(+) by high ionic strength, which activated the activity of the wild-type enzyme. These results suggest a critical role of Tyr-180 in the catalytic function of this class of enzymes, in addition to functions of His-79 in the coenzyme binding and chemical steps of the reaction.  相似文献   

2.
3.
Aminopropionaldehyde dehydrogenase was purified to apparent homogeneity from 1,3-diaminopropane-grown cells of Arthrobacter sp. TMP-1. The native molecular mass and the subunit molecular mass of the enzyme were approximately 20,5000 and 52,000, respectively, suggesting that the enzyme is a tetramer of identical subunits. The apparent Michaelis constant (K(m)) for 1,3-diaminopropane was approximately 3 microM. The enzyme equally used both NAD(+) and NADP(+) as coenzymes. The apparent K(m) values for NAD(+) and NADP(+) were 255 microM and 108 microM, respectively. The maximum reaction rates (V(max)) for NAD(+) and NADP(+) were 102 and 83.3 micromol min(-1) mg(-1), respectively. Some tested aliphatic aldehydes and aromatic aldehydes were inert as substrates. The optimum pH was 8.0-8.5. The enzyme was sensitive to sulfhydryl group-modifying reagents.  相似文献   

4.
AIM: A novel NADP(+)-dependent L-1-amino-2-propanol dehydrogenase was isolated from Rhodococcus erythropolis MAK154, and characterized. METHODS AND RESULTS: The enzyme was inducibly produced on cultivation with aminoalcohols such as 1-amino-2-propanol, 1-amino-2-butanol and 2-aminocyclohexanol. The enzyme catalyses the NADP(+)-dependent oxidation of several aminoalcohols, and also the NADPH-dependent asymmetric reduction of an aminoketone compound to a double chiral aminoalcohol, d-pseudoephedrine. Amino acid sequence analysis showed that the enzyme might belong to the short-chain dehydrogenase/reductase family. CONCLUSIONS: NADP(+)-dependent L-1-amino-2-propanol dehydrogenase isolated from R. erythropolis MAK154 reversibly catalysed dehydrogenation of aminoalcohols, and exhibited a unique sterospecifity for the reduction reaction. SIGNIFICANCE AND IMPACT OF THE STUDY: The enzyme is a promising catalyst for the production of double chiral compound, d-pseudoephedrine, from prochiral substrate.  相似文献   

5.
This study is concerned with the development and application of kinetic locking-on and auxiliary tactics for bioaffinity purification of NADP(+)-dependent dehydrogenases, specifically (1) the synthesis and characterization of highly substituted N(6)-linked immobilized NADP(+) derivatives using a rapid solid-phase modular approach; (2) the evaluation of the N(6)-linked immobilized NADP(+) derivatives for use with the kinetic locking-on strategy for bioaffinity purification of NADP(+)-dependent dehydrogenases: Model bioaffinity chromatographic studies with glutamate dehydrogenase from bovine liver (GDH with dual cofactor specificity, EC 1.4.1.3) and glutamate dehydrogenase from Candida utilis (GDH which is NADP(+)-specific, EC 1.4.1.4); (3) the selection of an effective "stripping ligand" for NADP(+)-dehydrogenase bioaffinity purifications using N(6)-linked immobilized NADP(+) derivatives in the locking-on mode; and (4) the application of the developed bioaffinity chromatographic system to the purification of C. utilis GDH from a crude cellular extract.Results confirm that the newly developed N(6)-linked immobilized NADP(+) derivatives are suitable for the one-step bioaffinity purification of NADP(+)-dependent GDH provided that they are used in the locking-on mode, steps are taken to inhibit alkaline phosphatase activity in crude cellular extracts, and 2',5'-ADP is used as the stripping ligand during chromatography. The general principles described here are supported by a specific sample enzyme purification; the purification of C. utilis GDH to electrophoretic homogeneity in a single bioaffinity chromatographic step (specific activity, 9.12 micromol/min/mg; purification factor, 83.7; yield 88%). The potential for development of analogous bioaffinity systems for other NADP(+)-dependent dehydrogenases is also discussed.  相似文献   

6.
NAD(P)H-dependent d-xylose reductase is a homodimeric oxidoreductase that belongs to the aldo-keto reductase superfamily. The enzyme has the special function to catalyze the first step in the assimilation of xylose into yeast metabolic pathways. Performing this function via reducing the open chain xylose to xylitol, the xylose reductase of Pichia stipitis is one of the most important enzymes that can be used to construct recombinant Saccharomyces cerevisiae strain for utilizing xylose and producing alcohol. To investigate into the interaction mechanism of the enzyme with its ligand NAD and NADP, the 3D structure was developed for the NAD(P)H-dependent d-xylose reductase from P. stipitis. With the 3D structure, the molecular docking operations were conducted to find the most stable bindings of the enzyme with NAD and NADP, respectively. Based on these results, the binding pockets of the enzyme for NAD and NADP have been explicitly defined. It has been found that the residues in forming the binding pockets for both NAD and NADP are almost the same and mainly hydrophilic. These findings may be used to guide mutagenesis studies, providing useful clues to modify the enzyme to improve the utilization of xylose for producing alcohol. Also, because human aldose reductases have the function to reduce the open chain form of glucose to sorbitol, a process physiologically significant for diabetic patients at the time that their blood glucose levels are elevated, the information gained through this study may also stimulate the development of new strategies for therapeutic treatment of diabetes.  相似文献   

7.
An L-galactonate dehydratase and the corresponding gene were identified from the mould Hypocrea jecorina (Trichoderma reesei). This novel enzyme converts L-galactonate to L-threo-3-deoxy-hexulosonate (2-keto-3-deoxy-L-galactonate). The enzyme is part of the fungal pathway for D-galacturonic acid catabolism, a pathway which is only partly known. It is the second enzyme of this pathway after the D-galacturonic acid reductase. L-galactonate dehydratase activity is present in H. jecorina cells grown on D-galacturonic acid but absent when other carbon sources are used for growth. A deletion of the L-galactonate dehydratase gene in H. jecorina results in a strain with no growth on D-galacturonic acid. The active enzyme was produced in the heterologous host Saccharomyces cerevisiae and characterized. It exhibited activity with L-galactonate and D-arabonate where the hydroxyl group of the C2 is in L- and the hydroxyl group of the C3 is in D-configuration in the Fischer projection. However, it did not exhibit activity with D-galactonate, D-gluconate, L-gulonate or D-xylonate where the hydroxyl groups of the C2 and C3 are in different configuration.  相似文献   

8.
Arabinoxylan (`Cellace') corn fiber, containing 28.1% (w/w) as l-arabinose and 32.8% (w/w) as d-xylose, was hydrolyzed by a crude enzyme containing -xylanase, -xylosidase and -l-arabinofuranosidase originating from the extracellular culture broth of Penicillium funiculosum. The resultant hydrolysate contained l-arabinose, d-xylose and small amounts of other mono- and oligosaccharides. The l-arabinose and d-xylose were 21.3% (w/w) and 18.7% (w/w), respectively, based on the initial arabinoxylan. Williopsis saturnus var. saturnus, which metabolizes d-xylose without using l-arabinose, was aerobically cultured in the hydrolysate at 30 °C for 96 h. The sugar solution after removal of yeast cells contained l-arabinose and d-xylose which were 20.3% (w/w) and 0.002% (w/w), respectively, of the initial arabinoxylan. The solution was decolorized with activated carbon, and deionized with cation- and anion-exchange resins. The clear sugar solution thus obtained was composed of l-arabinose and d-xylose which were 19.3% (w/w) and 0.001% (w/w), respectively, of the initial arabinoxylan. After concentration of the sugar solution under reduced pressure, followed by crystallization of l-arabinose, 16% (w/w) l-arabinose (based on the initial arabinoxylan) was obtained as crude crystals. No d-xylose was detected in the final preparation.  相似文献   

9.
A novel thermostable NAD(P)H oxidase from the hyperthermophilic archaeon Thermococcus kodakarensis KOD1 (TkNOX) catalyzes oxidation of NADH and NADPH with oxygen from atmospheric air as an electron acceptor. Although the optimal temperature of TkNOX is >90°C, it also shows activity at 30°C. This enzyme was used for the regeneration of both NADP(+) and NAD(+) in alcohol dehydrogenase (ADH)-catalyzed enantioselective oxidation of racemic 1-phenylethanol. NADP(+) regeneration at 30°C was performed by TkNOX coupled with (R)-specific ADH from Lactobacillus kefir, resulting in successful acquisition of optically pure (S)-1-phenylethanol. The use of TkNOX with moderately thermostable (S)-specific ADH from Rhodococcus erythropolis enabled us to operate the enantioselective bioconversion accompanying NAD(+) regeneration at high temperatures. Optically pure (R)-1-phenylethanol was successfully obtained by this system after a shorter reaction time at 45-60°C than that at 30°C, demonstrating an advantage of the combination of thermostable enzymes. The ability of TkNOX to oxidize both NADH and NADPH with remarkable thermostability renders this enzyme a versatile tool for regeneration of the oxidized nicotinamide cofactors without the need for extra substrates other than dissolved oxygen from air.  相似文献   

10.
ADP-L-glycero-D-mannoheptose 6-epimerase is required for lipopolysaccharide inner core biosynthesis in several genera of Gram-negative bacteria. The enzyme contains both fingerprint sequences Gly-X-Gly-X-X-Gly and Gly-X-X-Gly-X-X-Gly near its N terminus, which is indicative of an ADP binding fold. Previous studies of this ADP-l-glycero-D-mannoheptose 6-epimerase (ADP-hep 6-epimerase) were consistent with an NAD(+) cofactor. However, the crystal structure of this ADP-hep 6-epimerase showed bound NADP (Deacon, A. M., Ni, Y. S., Coleman, W. G., Jr., and Ealick, S. E. (2000) Structure 5, 453-462). In present studies, apo-ADP-hep 6-epimerase was reconstituted with NAD(+), NADP(+), and FAD. In this report we provide data that shows NAD(+) and NADP(+) both restored enzymatic activity, but FAD could not. Furthermore, ADP-hep 6-epimerase exhibited a preference for binding of NADP(+) over NAD(+). The K(d) value for NADP(+) was 26 microm whereas that for NAD(+) was 45 microm. Ultraviolet circular dichroism spectra showed that apo-ADP-hep 6-epimerase reconstituted with NADP(+) had more secondary structure than apo-ADP-hep 6-epimerase reconstituted with NAD(+). Perchloric acid extracts of the purified enzyme were assayed with NAD(+)-specific alcohol dehydrogenase and NADP(+)-specific isocitric dehydrogenase. A sample of the same perchloric acid extract was analyzed in chromatographic studies, which demonstrated that ADP-hep 6-epimerase binds NADP(+) in vivo. A structural comparison of ADP-hep 6-epimerase with UDP-galactose 4-epimerase, which utilizes an NAD(+) cofactor, has identified the regions of ADP-hep 6-epimerase, which defines its specificity for NADP(+).  相似文献   

11.
A d-galacturonic acid reductase and the corresponding gene were identified from the mold Hypocrea jecorina (Trichoderma reesei). We hypothesize that the enzyme is part of a fungal d-galacturonic acid catabolic pathway which has not been described previously and which is distinctly different from the bacterial pathway. H. jecorina grown on d-galacturonic acid exhibits an NADPH-dependent d-galacturonic acid reductase activity. This activity is absent when the mold is grown on other carbon sources. The d-galacturonic acid reductase was purified, and tryptic digests of the purified protein were sequenced. The open reading frame of the corresponding gene was then cloned from a cDNA library. The open reading frame was functionally expressed in the yeast Saccharomyces cerevisiae. A histidine-tagged protein was purified, and the enzyme kinetics were characterized. The enzyme converts in a reversible reaction from d-galacturonic acid and NADPH to l-galactonic acid and NADP. The enzyme also exhibits activity with d-glucuronic acid and dl-glyceraldehyde.  相似文献   

12.
1. Kinetic experiments suggested the possible existence of at least two different NAD(+)-dependent aldehyde dehydrogenases in rat liver. Distribution studies showed that one enzyme, designated enzyme I, was exclusively localized in the mitochondria and that another enzyme, designated enzyme II, was localized in both the mitochondria and the microsomal fraction. 2. A NADP(+)-dependent enzyme was also found in the mitochondria and the microsomal fraction and it is suggested that this enzyme is identical with enzyme II. 3. The K(m) for acetaldehyde was apparently less than 10mum for enzyme I and 0.9-1.7mm for enzyme II. The K(m) for NAD(+) was similar for both enzymes (20-30mum). The K(m) for NADP(+) was 2-3mm and for acetaldehyde 0.5-0.7mm for the NADP(+)-dependent activity. 4. The NAD(+)-dependent enzymes show pH optima between 9 and 10. The highest activity was found in pyrophosphate buffer for both enzymes. In phosphate buffer there was a striking difference in activity between the two enzymes. Compared with the activity in pyrophosphate buffer, the activity of enzyme II was uninfluenced, whereas the activity of enzyme I was very low. 5. The results are compared with those of earlier investigations on the distribution of aldehyde dehydrogenase and with the results from purified enzymes from different sources.  相似文献   

13.
An NADP(+)-dependent alcohol dehydrogenase was found in Euglena gracilis Z grown on 1-hexanol, while it was detected at low activity in cells grown on ethanol or glucose as a carbon source, indicating that the enzyme is induced by the addition of 1-hexanol into the medium as a carbon source. This enzyme was extremely unstable, even at 4 degrees C, unless 20% ethylene glycol was added. The optimal pH was 8.8-9.0 for oxidation reaction. The apparent K(m) values for 1-hexanol and NADP(+) were found to be 6.79 mM and 46.7 microM for this enzyme, respectively. The substrate specificity of this enzyme was very different from that of already purified NAD(+)-specific ethanol dehydrogenase by showing the highest activity with 1-hexanol as a substrate, followed by 1-pentanol and 1-butanol, and there was very little activity with ethanol and 1-propanol. This enzyme was active towards the primary alcohols but not secondary alcohols. Accordingly, since the NADP(+)-specific enzyme was separated on DEAE cellulose column, Euglena was confirmed to contain a novel enzyme to be active towards middle and long-chain length of fatty alcohols.  相似文献   

14.
For better understanding of the coenzyme specificity in NAD-dependent MDH (tMDH) from Thermus flavus AT-62, we determined the crystal structures of tMDH-NADP(H) complex at maximally 1.65 A resolution. The overall structure is almost the same as that of the tMDH-NADH complex. However, NADP(H) binds to tMDH in the reverse orientation, where adenine occupies the position near the catalytic center and nicotinamide is positioned at the adenine binding site of the tMDH-NADH complex. Consistent with this, kinetic analysis of the malate-oxidizing reaction revealed that NADP(+) inhibited tMDH at high concentrations. This has provided the first evidence for the alternative binding mode of the nicotinamide coenzyme, that has pseudo-symmetry in its structure, in a single enzyme.  相似文献   

15.
NADP(+)-isocitrate dehydrogenase (NADP(+)-IDH) from the dinitrogen-fixing filamentous cyanobacterium Anabaena sp. strain PCC 7120 was purified to homogeneity. The native enzyme is composed of two identical subunits (M(r), 57,000) and cross-reacts with antibodies obtained against the previously purified NADP(+)-IDH from the unicellular cyanobacterium Synechocystis sp. strain PCC 6803. Anabaena NADP(+)-IDH resembles in its physicochemical and kinetic parameters the typical dimeric IDHs from prokaryotes. The gene encoding Anabaena NADP(+)-IDH was cloned by complementation of an Escherichia coli icd mutant with an Anabaena genomic library. The complementing DNA was located on a 6-kb fragment. It encodes an NADP(+)-IDH that has the same mobility as that of Anabaena NADP(+)-IDH on nondenaturing polyacrylamide gels. The icd gene was subcloned and sequenced. Translation of the nucleotide sequence gave a polypeptide of 473 amino acids that showed high sequence similarity to the E. coli enzyme (59% identity) and with IDH1 and IDH2, the two subunits of the heteromultimeric NAD(+)-IDH from Saccharomyces cerevisiae (30 to 35% identity); however, a low level of similarity to NADP(+)-IDHs of eukaryotic origin was found (23% identity). Furthermore, Anabaena NADP(+)-IDH contains a 44-residue amino acid sequence in its central region that is absent in the other IDHs so far sequenced. Attempts to generate icd mutants by insertional mutagenesis were unsuccessful, suggesting an essential role of IDH in Anabaena sp. strain PCC 7120.  相似文献   

16.
L Zhang  B Ahvazi  R Szittner  A Vrielink  E Meighen 《Biochemistry》1999,38(35):11440-11447
The fatty aldehyde dehydrogenase from the luminescent bacterium, Vibrio harveyi (Vh-ALDH), is unique with respect to its high specificity for NADP(+) over NAD(+). By mutation of a single threonine residue (Thr175) immediately downstream of the beta(B) strand in the Rossmann fold, the nucleotide specificity of Vh-ALDH has been changed from NADP(+) to NAD(+). Replacement of Thr175 by a negatively charged residue (Asp or Glu) resulted in an increase in k(cat)/K(m) for NAD(+) relative to that for NADP(+) of up to 5000-fold due to a decrease for NAD(+) and an increase for NADP(+) in their respective Michaelis constants (K(a)). Differential protection by NAD(+) and NADP(+) against thermal inactivation and comparison of the dissociation constants of NMN, 2'-AMP, 2'5'-ADP, and 5'-AMP for these mutants and the wild-type enzyme clearly support the change in nucleotide specificity. Moreover, replacement of Thr175 with polar residues (N, S, or Q) demonstrated that a more efficient NAD(+)-dependent enzyme T175Q could be created without loss of NADP(+)-dependent activity. Analysis of the three-dimensional structure of Vh-ALDH with bound NADP(+) showed that the hydroxyl group of Thr175 forms a hydrogen bond to the 2'-phosphate of NADP(+). Replacement with glutamic acid or glutamine strengthened interactions with NAD(+) and indicated why threonine would be the preferred polar residue at the nucleotide recognition site in NADP(+)-specific aldehyde dehydrogenases. These results have shown that the size and the structure of the residue at the nucleotide recognition site play the key roles in differentiating between NAD(+) and NADP(+) interactions while the presence of a negative charge is responsible for the decrease in interactions with NADP(+) in Vh-ALDH.  相似文献   

17.
Three alcohol dehydrogenases have been identified in Acinetobacter calcoaceticus sp. strain HO1-N: an NAD(+)-dependent enzyme and two NADP(+)-dependent enzymes. One of the NADP(+)-dependent alcohol dehydrogenases was partially purified and was specific for long-chain substrates. With tetradecanol as substrate an apparent Km value of 5.2 microM was calculated. This enzyme has a pI of 4.5 and a molecular mass of 144 kDa. All three alcohol dehydrogenases were constitutively expressed. Three aldehyde dehydrogenases were also identified: an NAD(+)-dependent enzyme, an NADP(+)-dependent enzyme and one which was nucleotide independent. The NAD(+)-dependent enzyme represented only 2% of the total activity and was not studied further. The NADP(+)-dependent enzyme was strongly induced by growth of cells on alkanes and was associated with hydrocarbon vesicles. With tetradecanal as substrate an apparent Km value of 0.2 microM was calculated. The nucleotide-independent aldehyde dehydrogenase could use either Würster's Blue or phenazine methosulphate (PMS) as an artificial electron acceptor. This enzyme represents approximately 80% of the total long-chain aldehyde oxidizing activity within the cell when the enzymes were induced by growing the cells on hexadecane. It is particulate but can be solubilized using Triton X-100. The enzyme has an apparent Km of 0.36 mM for decanal.  相似文献   

18.
Proton-translocating nicotinamide nucleotide transhydrogenases contain an NAD(H)-binding domain (dI), an NADP(H)-binding domain (dIII) and a membrane domain (dII) with the proton channel. Separately expressed and isolated dIII contains tightly bound NADP(H), predominantly in the oxidized form, possibly representing a so-called "occluded" intermediary state of the reaction cycle of the intact enzyme. Despite a K(d) in the micromolar to nanomolar range, this NADP(H) exchanges significantly with the bulk medium. Dissociated NADP(+) is thus accessible to added enzymes, such as NADP-isocitrate dehydrogenase, and can be reduced to NADPH. In the present investigation, dissociated NADP(H) was digested with alkaline phosphatase, removing the 2'-phosphate and generating NAD(H). Surprisingly, in the presence of dI, the resulting NADP(H)-free dIII catalyzed a rapid reduction of 3-acetylpyridine-NAD(+) by NADH, indicating that 3-acetylpyridine-NAD(+) and/or NADH interacts unspecifically with the NADP(H)-binding site. The corresponding reaction in the intact enzyme is not associated with proton pumping. It is concluded that there is a 2'-phosphate-binding region in dIII that controls tight binding of NADP(H) to dIII, which is not a required for fast hydride transfer. It is likely that this region is the Lys424-Arg425-Ser426 sequence and loops D and E. Further, in the intact enzyme, it is proposed that the same region/loops may be involved in the regulation of NADP(H) binding by an electrochemical proton gradent.  相似文献   

19.
Reduction and oxidation of steroids in the human gut are catalyzed by hydroxysteroid dehydrogenases of microorganisms. For the production of 12-ketochenodeoxycholic acid (12-Keto-CDCA) from cholic acid the biocatalytic application of the 12α-hydroxysteroid dehydrogenase of Clostridium group P, strain C 48-50 (HSDH) is an alternative to chemical synthesis. However, due to the intensive costs the necessary cofactor (NADP(+) ) has to be regenerated. The alcohol dehydrogenase of Thermoanaerobacter ethanolicus (ADH-TE) was applied to catalyze the reduction of acetone while regenerating NADP(+) . A mechanistic kinetic model was developed for the process development of cholic acid oxidation using HSDH and ADH-TE. The process model was derived by identifying the parameters for both enzymatic models separately using progress curve measurements of batch processes over a broad range of concentrations and considering the underlying ordered bi-bi mechanism. Both independently derived kinetic models were coupled via mass balances to predict the production of 12-Keto-CDCA with HSDH and integrated cofactor regeneration with ADH-TE and acetone as co-substrate. The prediction of the derived model was suitable to describe the dynamics of the preparative 12-Keto-CDCA batch production with different initial reactant and enzyme concentrations. These datasets were used again for parameter identification. This led to a combined model which excellently described the reaction dynamics of biocatalytic batch processes over broad concentration ranges. Based on the identified process model batch process optimization was successfully performed in silico to minimize enzyme costs. By using 0.1 mM NADP(+) the HSDH concentration can be reduced to 3-4 μM and the ADH concentration to 0.4-0.6 μM to reach the maximal possible conversion of 100 mM cholic acid within 48 h. In conclusion, the identified mechanistic model offers a powerful tool for a cost-efficient process design.  相似文献   

20.
In this study the ability of various sugars and sugar alcohols to induce aldose reductase (xylose reductase) and xylitol dehydrogenase (xylulose reductase) activities in the yeast Candida tenuis was investigated. Both enzyme activities were induced when the organism was grown on d-xylose or l-arabinose as well as on the structurally related sugars d-arabinose or d-lyxose. Mixtures of d-xylose with the more rapidly metabolizable sugar d-glucose resulted in a decrease in the levels of both enzymes formed. These results show that the utilization of d-xylose by C. tenuis is regulated by induction and catabolite repression. Furthermore, the different patterns of induction on distinct sugars suggest that the synthesis of both enzymes is not under coordinate control.  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