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.
The fungal L-arabinose pathway consists of five enzymes, aldose reductase, L-arabinitol 4-dehydrogenase, L-xylulose reductase, xylitol dehydrogenase, and xylulokinase. All the genes encoding the enzymes of this pathway are known except for that of L-xylulose reductase (EC 1.1.1.10). We identified a gene encoding this enzyme from the filamentous fungus Trichoderma reesei (Hypocrea jecorina). The gene was named lxr1. It was overexpressed in the yeast Saccharomyces cerevisiae, and the enzyme activity was confirmed in a yeast cell extract. Overexpression of all enzymes of the L-arabinose pathway in S. cerevisiae led to growth of S. cerevisiae on L-arabinose; i.e., we could show that the pathway is active in a heterologous host. The lxr1 gene encoded a protein with 266 amino acids and a calculated molecular mass of 28 428 Da. The LXRI protein is an NADPH-specific reductase. It has activity with L-xylulose, D-xylulose, D-fructose, and L-sorbose. The highest affinity is toward L-xylulose (K(m) = 16 mM). In the reverse direction, we found activity with xylitol, D-arabinitol, D-mannitol, and D-sorbitol. It requires a bivalent cation for activity. It belongs to the protein family of short chain dehydrogenases. The enzyme is catalytically similar and homologous in sequence to a D-mannitol:NADP 2-dehydrogenase (EC 1.1.1.138).  相似文献   

2.
A polyol dehydrogenase was detected in cell extracts of the facultative phototrophic bacterium Rhodobacter sphaeroides strain Si 4 grown on D-glucitol (sorbitol) as the sole carbon source. The enzyme was purified 150-fold to apparent homogeneity by steps involving fractionated (NH4)2SO4 precipitation, chromatography on Q-Sepharose and phenyl-Sepharose, and FPLC on Superose 12. The relative molecular mass (Mr) of the native polyol dehydrogenase was 47,200 as calculated from its Stokes' radius (rs = 2.76 nm) and sedimentation coefficient (s20, w = 4.15 S). SDS/PAGE resulted in one single band representing a polypeptide with a Mr of 52,200, indicating that the native protein is a monomer. The isoelectric point of the polyol dehydrogenase was determined to be pH 4.3. The enzyme was specific for NAD+ and oxidized both D-glucitol and D-mannitol to D-fructose, as well as D-arabinitol to D-ribulose. The pH optimum of substrate oxidation was pH 9.0 in 0.1 M Tris/HCl and that of substrate reduction was pH 6.5 in 0.1 M potassium phosphate. The reactions exhibited normal Michaelis-Menten kinetics allowing the estimation of KM values for NAD+ (0.18 mM) in the presence of D-glucitol, and for D-glucitol (31.8 mM), D-mannitol (0.29 mM) and D-arabinitol (1.8 mM), respectively. The KM value for D-fructose was 16.3 mM and that for NADH 0.02 mM. The equilibrium constants determined for the conversion of D-mannitol, D-glucitol and D-arabinitol were 4.5 nM, 0.58 nM and 80 pM, respectively. Based on the catalytic preference of the polyol dehydrogenase for D-mannitol, an enzymatic assay for D-mannitol was elaborated.  相似文献   

3.
Mannitol dehydrogenase, NADP specific (EC 1.1.1.138), was purified from mycelium of Aspergillus parasiticus (1-11-105 Whl). The enzyme had a molecular weight of 1.4 X 10(5) and was composed of four subunits of apparently equal size. The substrate specificity was limited to D-mannitol, D-glucitol, D-arabinitol, 1-deoxy-D-mannitol, and 1-deoxy-D-glucitol. Zinc ion was a powerful inhibitor of the enzyme, inhibition being competitive with respect to mannitol, with Ki and 1 microM. It is proposed that the stimulation of polyketide synthesis by zinc ion may be mediated in part by inhibition of mannitol dehydrogenase.  相似文献   

4.
M Slatner  B Nidetzky  K D Kulbe 《Biochemistry》1999,38(32):10489-10498
To characterize catalysis by NAD-dependent long-chain mannitol 2-dehydrogenases (MDHs), the recombinant wild-type MDH from Pseudomonas fluorescens was overexpressed in Escherichia coli and purified. The enzyme is a functional monomer of 54 kDa, which does not contain Zn(2+) and has B-type stereospecificity with respect to hydride transfer from NADH. Analysis of initial velocity patterns together with product and substrate inhibition patterns and comparison of primary deuterium isotope effects on the apparent kinetic parameters, (D)k(cat), (D)(k(cat)/K(NADH)), and (D)(k(cat)/K(fructose)), show that MDH has an ordered kinetic mechanism at pH 8.2 in which NADH adds before D-fructose, and D-mannitol and NAD are released in that order. Isomerization of E-NAD to a form which interacts with D-mannitol nonproductively or dissociation of NAD from the binary complex after isomerization is the slowest step (>/=110 s(-)(1)) in D-fructose reduction at pH 8.2. Release of NADH from E-NADH (32 s(-)(1)) is the major rate-limiting step in mannitol oxidation at this pH. At the pH optimum for D-fructose reduction (pH 7.0), the rate of hydride transfer contributes significantly to rate limitation of the catalytic cascade and the overall reaction. (D)(k(cat)/K(fructose)) decreases from 2.57 at pH 7.0 to a value of 相似文献   

5.
A polyol dehydrogenase of broad specificity was purified 178-fold from extracts of the filamentous fungus Cephalosporium chrysogenum. The enzyme was found to act as an oxido-reductase in two substrate-coenzyme systems: D-sorbitol (or xylitol)-nicotinamide-adenine dinucleotide (NAD) and D-mannitol-nicotinamide adenine dinucleotide phosphate (NADP). The dehydrogenase was composed of five isozymes, which, as a mixture, exhibited these properties: Km to D-sorbitol and D-mannitol, 7.15 to 10(-2) M; PH optimum, 9 to 10; molecular weight, 300,000 subunit weight, 29,000; PI, 5.8 to 7.5. The NADP-linked activity was labile to treatment with heat or ethylenediaminetetraacetic acid. Mixed substrate assays support the hypothesis that both NAD-, and NADP-linked activities are associated with isozymes of a single dehydrogenase.  相似文献   

6.
ABSTRACT

Gluconobacter frateurii CHM 43 have D-mannitol dehydrogenase (quinoprotein glycerol dehydrogenase) and flavoprotein D-fructose dehydrogenase in the membranes. When the two enzymes are functional, D-mannitol is converted to 5-keto-D-fructose with 65% yield when cultivated on D-mannitol. 5-Keto-D-fructose production with almost 100% yield was realized with the resting cells. The method proposed here should give a smart strategy for 5-keto-D-fructose production.  相似文献   

7.
The D-sorbitol dehydrogenase gene, sldA, and an upstream gene, sldB, encoding a hydrophobic polypeptide, SldB, of Gluconobacter suboxydans IFO 3255 were disrupted in a check of their biological functions. The bacterial cells with the sldA gene disrupted did not produce L-sorbose by oxidation of D-sorbitol in resting-cell reactions at pHs 4.5 and 7.0, indicating that the dehydrogenase was the main D-sorbitol-oxidizing enzyme in this bacterium. The cells did not produce D-fructose from D-mannitol or dihydroxyacetone from glycerol. The disruption of the sldB gene resulted in undetectable oxidation of D-sorbitol, D-mannitol, or glycerol, although the cells produced the dehydrogenase. The cells with the sldB gene disrupted produced more of what might be signal-unprocessed SldA than the wild-type cells did. SldB may be a chaperone-like component that assists signal processing and folding of the SldA polypeptide to form active D-sorbitol dehydrogenase.  相似文献   

8.
An enzymatic, NAD(H)-dependent process for the efficient production of D-mannitol from D-fructose as one single product is described and optimized with respect to productivity at high substrate conversion. Stereospecific reduction of D-fructose is catalyzed by recombinant mannitol dehydrogenase from Pseudomonas fluorescens DSM 50106, overexpressed in Escherichia coli. Regeneration of NADH is accomplished by formate dehydrogenase-mediated oxidation of formate into CO2, thus avoiding byproduct formation and yielding total turnover numbers for the coenzyme of approximately 1000 for a single round of D-fructose conversion. In optimized batchwise reduction of D-fructose, a D-mannitol productivity of 2.25 g/(L h) was obtained for a final product concentration of 72g/L and a D-fructose conversion of 80%. D-Mannitol was crystallized from the ultrafiltered product solution in 97% purity and 85% recovery, thus also allowing reuse of enzymes for repeated batchwise production of D-mann!itol!.  相似文献   

9.
An in vivo system was developed for the biotransformation of D-fructose into D-mannitol by the expression of the gene mdh encoding mannitol dehydrogenase (MDH) from Leuconostoc pseudomesenteroides ATCC12291 in Bacillus megaterium. The NADH reduction equivalents necessary for MDH activity were regenerated via the oxidation of formate to carbon dioxide by coexpression of the gene fdh encoding Mycobacterium vaccae N10 formate dehydrogenase (FDH). High-level protein production of MDH in B. megaterium required the adaptation of the corresponding ribosome binding site. The fdh gene was adapted to B. megaterium codon usage via complete chemical gene synthesis. Recombinant B. megaterium produced up to 10.60 g/L D-mannitol at the shaking flask scale. Whole cell biotransformation in a fed-batch bioreactor increased D-mannitol concentration to 22.00 g/L at a specific productivity of 0.32 g D-mannitol (gram cell dry weight)(-1) h(-1) and a D-mannitol yield of 0.91 mol/mol. The nicotinamide adenine dinucleotide (NAD(H)) pool of the B. megaterium producing D-mannitol remained stable during biotransformation. Intra- and extracellular pH adjusted itself to a value of 6.5 and remained constant during the process. Data integration revealed that substrate uptake was the limiting factor of the overall biotransformation. The information obtained identified B. megaterium as a useful production host for D-mannitol using a resting cell biotransformation approach.  相似文献   

10.
Clostridial glutamate dehydrogenase mutants, designed to accommodate the 2'-phosphate of disfavoured NADPH, showed the expected large specificity shifts with NAD(P)H. Puzzlingly, similar assays with oxidized cofactors initially revealed little improvement with NADP(+) , although rates with NAD(+) were markedly diminished. This article reveals that the enzyme's discrimination in favour of NAD(+) and against NADP(+) had been greatly underestimated and has indeed been abated by a factor of >?16,000 by the mutagenesis. Initially, stopped-flow studies of the wild-type enzyme showed a burst increase of A(340) with NADP(+) but not NAD(+), with amplitude depending on the concentration of the coenzyme, rather than enzyme. Amplitude also varied with the commercial source of the NADP(+). FPLC, HPLC and mass spectrometry identified NAD(+) contamination ranging from 0.04 to 0.37% in different commercial samples. It is now clear that apparent rates of NADP(+) utilization mainly reflected the reduction of contaminating NAD(+), creating an entirely false view of the initial coenzyme specificity and also of the effects of mutagenesis. Purification of the NADP(+) eliminated the burst. With freshly purified NADP(+), the NAD(+) : NADP(+) activity ratio under standard conditions, previously estimated as 300 : 1, is 11,000. The catalytic efficiency ratio is even higher at 80,000. Retested with pure cofactor, mutants showed marked specificity shifts in the expected direction, for example, 16 200 fold change in catalytic efficiency ratio for the mutant F238S/P262S, confirming that the key structural determinants of specificity have been successfully identified. Of wider significance, these results underline that, without purification, even the best commercial coenzyme preparations are inadequate for such studies.  相似文献   

11.
NADP(+)-dependent D-threonine dehydrogenase (EC 1.1.1.-), which catalyzes the oxidation of the 3-hydroxyl group of D-threonine, was purified to homogeneity from a crude extract of Pseudomonas cruciviae IFO 12047. The enzyme had a molecular mass of about 60,000 Da and consisted of two identical subunits. In addition to D-threonine, D-threo-3-phenylserine, D-threo-3-thienylserine, and D-threo-3-hydroxynorvaline were also substrates. However, the other isomers of threonine and 3-phenylserine were inert. The enzyme showed maximal activity at pH 10.5 for the oxidation of D-threonine. The enzyme required NADP+. NAD+ showed only slight activity. The enzyme was not inhibited by EDTA, o-phenanthroline, alpha,alpha'-dipyridyl, HgCl2, or p-chloromercuribenzoate but was inhibited by tartronate, malonate, pyruvate, and DL-2-hydroxybutyrate. The inhibition by these organic acids was competitive against D-threonine. Initial-velocity and product inhibition studies suggested that the oxidation proceeded through a sequential ordered Bi Bi mechanism. The Michaelis constants for D-threonine and NADP+ were 13 and 0.12 mM, respectively.  相似文献   

12.
Gluconate:NADP 5-oxidoreductase (GNO) from the acetic acid bacterium Gluconobacter oxydans subsp. oxydans DSM3503 was purified to homogeneity. This enzyme is involved in the nonphosphorylative, ketogenic oxidation of glucose and oxidizes gluconate to 5-ketogluconate. GNO was localized in the cytoplasm, had an isoelectric point of 4.3, and showed an apparent molecular weight of 75,000. In sodium dodecyl sulfate gel electrophoresis, a single band appeared corresponding to a molecular weight of 33,000, which indicated that the enzyme was composed of two identical subunits. The pH optimum of gluconate oxidation was pH 10, and apparent Km values were 20.6 mM for the substrate gluconate and 73 microM for the cosubstrate NADP. The enzyme was almost inactive with NAD as a cofactor and was very specific for the substrates gluconate and 5-ketogluconate. D-Glucose, D-sorbitol, and D-mannitol were not oxidized, and 2-ketogluconate and L-sorbose were not reduced. Only D-fructose was accepted, with a rate that was 10% of the rate of 5-ketogluconate reduction. The gno gene encoding GNO was identified by hybridization with a gene probe complementary to the DNA sequence encoding the first 20 N-terminal amino acids of the enzyme. The gno gene was cloned on a 3.4-kb DNA fragment and expressed in Escherichia coli. Sequencing of the gene revealed an open reading frame of 771 bp, encoding a protein of 257 amino acids with a predicted relative molecular mass of 27.3 kDa. Plasmid-encoded gno was functionally expressed, with 6.04 U/mg of cell-free protein in E. coli and with 6.80 U/mg of cell-free protein in G. oxydans, which corresponded to 85-fold overexpression of the G. oxydans wild-type GNO activity. Multiple sequence alignments showed that GNO was affiliated with the group II alcohol dehydrogenases, or short-chain dehydrogenases, which display a typical pattern of six strictly conserved amino acid residues.  相似文献   

13.
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   

14.
5-Ketofructose reductase [D(-)fructose:(NADP+) 5-oxidoreductase] was purified to homogeneity from Erwinia citreus and demonstrated to catalyse the reversible NADPH-dependent reduction of 5-ketofructose (D-threo-2,5-hexodiulose) to D-fructose. The enzyme appeared as a single species upon analyses by SDS/polyacrylamide-gel electrophoresis and isoelectric focusing with an apparent relative molecular mass of 40,000 and an isoelectric point of 4.4. The amino acid composition of the enzyme and the N-terminal sequence of the first 39 residues are described. The steady-state kinetic mechanism was an ordered one with NADPH binding first to the enzyme and then to 5-ketofructose, and the order of product release was D-fructose followed by NADP+. The reversible nature of the reaction offers the possibility of using this enzyme for the determination of D-fructose.  相似文献   

15.
Previously we found that replacement of seven amino acid residues in a loop region markedly shifted the coenzyme specificity of malate dehydrogenase from NAD(H) toward NADP(H). In the present study, we replaced the seven amino acid residues in the corresponding region of an NAD(H)-dependent lactate dehydrogenase with those of NADP(H)-dependent malate dehydrogenase, and examined the coenzyme specificity of the resulting mutant enzyme. Coenzyme specificity was significantly shifted by 399-fold toward NADPH when k cat/Km(coenzyme) was used as the measure of coenzyme specificity. The effect of the replacements on coenzyme specificity is discussed based on in silico simulation of the three-dimensional structure of the lactate dehydrogenase mutant.  相似文献   

16.
The kinetic mechanism of glucose dehydrogenase (EC 1.1.1.47) from Halobacterium salinarum was studied by initial velocity and product inhibition methods. The results suggest that both, in the forward and reverse direction, the reaction mechanism is of Bi Bi sequential ordered type involving formation of ternary complexes. NADP+ adds first and NADPH formed dissociates from the enzyme last. For the reverse direction, NADPH adds first and NADP+ leaves last. Product inhibition experiments indicate that (a), the coenzymes compete for the same site and form of the enzyme and (b), ternary abortive complexes of enzyme-NADP(+)-glucono-delta-lactone and enzyme-NADPH-glucose are formed. All the other inhibitions are noncompetitive.  相似文献   

17.
The human pathogenic fungus Aspergillus fumigatus accumulates large amounts of intracellular mannitol to enhance its resistance against defense strategies of the infected host. To explore their currently unknown roles in mannitol metabolism, we studied A. fumigatus mannitol-1-phosphate 5-dehydrogenase (AfM1PDH) and mannitol 2-dehydrogenase (AfM2DH), each recombinantly produced in Escherichia coli, and performed a detailed steady-state kinetic characterization of the two enzymes at 25 °C and pH 7.1. Primary kinetic isotope effects resulting from deuteration of alcohol substrate or NADH showed that, for AfM1PDH, binding of D-mannitol 1-phosphate and NAD(+) is random, whereas D-fructose 6-phosphate binds only after NADH has bound to the enzyme. Binding of substrate and NAD(H) by AfM2DH is random for both D-mannitol oxidation and D-fructose reduction. Hydride transfer is rate-determining for D-mannitol 1-phosphate oxidation by AfM1PDH (k(cat) = 10.6 s(-1)) as well as D-fructose reduction by AfM2DH (k(cat) = 94 s(-1)). Product release steps control the maximum rates in the other direction of the two enzymatic reactions. Free energy profiles for the enzymatic reaction under physiological boundary conditions suggest that AfM1PDH primarily functions as a D-fructose-6-phosphate reductase, whereas AfM2DH acts in D-mannitol oxidation, thus establishing distinct routes for production and mobilization of mannitol in A. fumigatus. ATP, ADP and AMP do not affect the activity of AfM1PDH, suggesting the absence of flux control by cellular energy charge at the level of D-fructose 6-phosphate reduction. AfM1PDH is remarkably resistant to inactivation by heat (half-life at 40 °C of 20 h), consistent with the idea that formation of mannitol is an essential component of the temperature stress response of A. fumigatus. Inhibition of AfM1PDH might be a useful target for therapy of A. fumigatus infections.  相似文献   

18.
Abstract A sorbitol dehydrogenase was purified from the membrane fraction of Gluconobacter suboxydans KCTC 2111 (= ATCC 621) by chromatography on CM-, DEAE-, Mono S and Superose 12 columns. The purified enzyme showed a single activity band upon nondenaturing polyacrylamide gel electrophoresis (PAGE) and three subunits of 75, 50 and 14 kDa upon SDS-PAGE. When purified preparations of the enzyme were reconstituted with pyrroloquinoline quinone (PQQ), the specific enzyme activity was significantly increased (up to 9-fold). The absorption spectrum of purified sorbitol dehydrogenase in the reduced state exhibited three absorption maxima (417, 522 and 552 nm) which is in accordance with the typical absorption spectrum of cytochrome c . The 50 kDa subunit appeared as a red band on unstained SDS-gels suggesting its identity as a cytochrome. Fluorescence spectra of extracts from purified sorbitol dehydrogenase showed an excitation maximum at 370 nm and an emission maximum at 465 nm, which conformed to those of authentic PQQ. The purified enzyme showed a rather broad substrate specificity with significant activity toward D-mannitol (68%) and D-ribitol (70%) as well as D-sorbitol (100%). The PQQ-dependent sorbitol dehydrogenase described in this study is clearly different from the FAD-dependent sorbitol dehydrogenase from G. suboxydans var. α IFO 3254 strain in its cofactor requirement and substrate specificity.  相似文献   

19.
A new NADP(H)-dependent alcohol dehydrogenase (the YCR105W gene product, ADHVII) has been identified in Saccharomyces cerevisiae. The enzyme has been purified to homogeneity and found to be a homodimer of 40 kDa subunits and a pI of 6.2-6.4. ADHVII shows a broad substrate specificity similar to the recently characterized ADHVI (64% identity), although they show some differences in kinetic properties. ADHVI and ADHVII are the only members of the cinnamyl alcohol dehydrogenase family in yeast. Simultaneous deletion of ADH6 and ADH7 was not lethal for the yeast. Both enzymes could participate in the synthesis of fusel alcohols, ligninolysis and NADP(H) homeostasis.  相似文献   

20.
F Fan  J A Lorenzen  B V Plapp 《Biochemistry》1991,30(26):6397-6401
In the three-dimensional structures of enzymes that bind NAD or FAD, there is an acidic residue that interacts with the 2'- and 3'-hydroxyl groups of the adenosine ribose of the coenzyme. The size and charge of the carboxylate might repel the binding of the 2'-phosphate group of NADP and explain the specificity for NAD. In the NAD-dependent alcohol dehydrogenases, Asp-223 (horse liver alcohol dehydrogenase sequence) appears to have this role. The homologous residue in yeast alcohol dehydrogenase I (residue 201 in the protein sequence) was substituted with Gly, and the D223G enzyme was expressed in yeast, purified, and characterized. The wild-type enzyme is specific for NAD. In contrast, the D223G enzyme bound and reduced NAD+ and NADP+ equally well, but, relative to wild-type enzyme, the dissociation constant for NAD+ was increased 17-fold, and the reactivity (V/K) on ethanol was decreased to 1%. Even though catalytic efficiency was reduced, yeast expressing the altered or wild-type enzyme grew at comparable rates, suggesting that equilibration of NAD and NADP pools is not lethal. Asp-223 participates in binding NAD and in excluding NADP, but it is not the only residue important for determining specificity for coenzyme.  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