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 32 毫秒
1.
2.
The washed cells of a gluconate-utilizing Corynebacterium strain grown in a gluconate- xylose medium produced xylitol from D-xylose in the presence of gluconate. The amount of xylitol was progressively increased with increasing gluconate concentration.

An extract of cells grown in the gluconate-xylose medium showed NADPH-dependent D-xylose reductase activity and NADP-dependent 6-phosphogluconate dehydrogenase activity.

These enzymes in the cell-free extract were purified by Sephadex G–100 gel filtration.

The reduction of D-xylose to xylitol was demonstrated by the coupling the D-xylose reductase activity to the 6-phosphogluconate dehydrogenase activity with NADP as a cofactor using the cell-free extract and the fractionated enzymes.  相似文献   

3.
Xylose reductase has been purified to apparent homogeneity from cell extracts of the fungus Cryptococcus flavus grown on D-xylose as carbon source. The enzyme, the first of its kind from the phylum Basidiomycota, is a functional dimer composed of identical subunits of 35.3 kDa mass and requires NADP(H) for activity. Steady-state kinetic parameters for the reaction, D-xylose + NADPH + H(+)<--> xylitol + NADP(+), have been obtained at pH 7.0 and 25 degrees C. The catalytic efficiency for reduction of D-xylose is 150 times that for oxidation of xylitol. This and the 3-fold tighter binding of NADPH than NADP(+) indicate that the enzyme is primed for unidirectional metabolic function in microbial physiology. Kinetic analysis of enzymic reduction of aldehyde substrates differing in hydrophobic and hydrogen bonding capabilities with binary enzyme-NADPH complex has been used to characterize the substrate-binding pocket of xylose reductase. Total transition state stabilization energy derived from bonding with non-reacting sugar hydroxyls is approximately 15 kJ/mol, with a major contribution of 5-8 kJ/mol made by interactions with the C-2(R) hydroxy group. The aldehyde binding site is approximately 1.2 times more hydrophobic than n-octanol and can accommodate linear alkyl chains of 相似文献   

4.
Mycobacterium smegmatis transformed D-xylulose to xylitol in washed cell reactions under aerobic and anaerobic conditions. The yield of xylitol reached about 70% in anaerobic conditions (in N2) by cells grown on media containing xylitol or D-mannitol. Cells immobilized with Ca-alginate had almost the same activity of xylitol production as washed cells.Xylitol was produced from D-xylose using commercial immobilized D-xylose isomerase from Bacillus coagulans and immobilized cells of M. smegmatis. From 10 g of D-xylose, 4 g of xylitol was produced and 5 g of D-xylose remained in the reaction mixture; no D-xylulose was detected.  相似文献   

5.
An NAD(+)-dependent D-xylose dehydrogenase, XylB, from Caulobacter crescentus was expressed in Saccharomyces cerevisiae, resulting in production of 17 ± 2 g D-xylonate l(-1) at 0.23 gl(-1)h(-1) from 23 g D-xylose l(-1) (with glucose and ethanol as co-substrates). D-Xylonate titre and production rate were increased and xylitol production decreased, compared to strains expressing genes encoding T. reesei or pig liver NADP(+)-dependent D-xylose dehydrogenases. D-Xylonate accumulated intracellularly to ~70 mgg(-1); xylitol to ~18 mgg(-1). The aldose reductase encoding gene GRE3 was deleted to reduce xylitol production. Cells expressing D-xylonolactone lactonase xylC from C. crescentus with xylB initially produced more extracellular D-xylonate than cells lacking xylC at both pH 5.5 and pH 3, and sustained higher production at pH 3. Cell vitality and viability decreased during D-xylonate production at pH 3.0. An industrial S. cerevisiae strain expressing xylB efficiently produced 43 g D-xylonate l(-1) from 49 g D-xylose l(-1).  相似文献   

6.
The compositions of intracellular pentose phosphate pathway enzymes have been examined in mutants of Pachysolen tannophilus NRRL Y-2460 which possessed enhanced D-xylose fermentation rates. The levels of oxidoreductive enzymes involved in converting D-xylose to D-xylulose via xylitol were 1.5–14.7-fold higher in mutants than in the parent. These enzymes were still under inductive control by D-xylose in the mutants. The D-xylose reductase activity (EC 1.1.1.21) which catalyses the conversion of D-xylose to xylitol was supported with either NADPH or NADH as coenzyme in all the mutant strains. Other enzyme specific activities that generally increased were: xylitol dehydrogenase (EC 1.1.1.9), 1.2–1.6-fold; glucose-6-phosphate dehydrogenase (EC 1.1.1.49), 1.9–2.6-fold; D-xylulose-5-phosphate phosphoketolase (EC 4.1.2.9), 1.2–2.6-fold; and alcohol dehydrogenase (EC 1.1.1.1), 1.5–2.7-fold. The increase of enzymatic activities, 5.3–10.3-fold, occurring in D-xylulokinase (EC 2.7.1.17), suggested a pivotal role for this enzyme in utilization of D-xylose by these mutants. The best ethanol-producing mutant showed the highest ratio of NADH- to NADPH-linked D-xylose reductase activity and high levels of all other pentose phosphate pathway enzymes assayed.  相似文献   

7.
Considerable interest in the D-xylose catabolic pathway of Pachysolen tannophilus has arisen from the discovery that this yeast is capable of fermenting D-xylose to ethanol. In this organism D-xylose appears to be catabolized through xylitol to D-xylulose. NADPH-linked D-xylose reductase is primarily responsible for the conversion of D-xylose to xylitol, while NAD-linked xylitol dehydrogenase is primarily responsible for the subsequent conversion of xylitol to D-xylulose. Both enzyme activities are readily detectable in cell-free extracts of P. tannophilus grown in medium containing D-xylose, L-arabinose, or D-galactose and appear to be inducible since extracts prepared from cells growth in media containing other carbon sources have only negligible activities, if any. Like D-xylose, L-arabinose and D-galactose were found to serve as substrates for NADPH-linked reactions in extracts of cells grown in medium containing D-xylose, L-arabinose, or D-galactose. These L-arabinose and D-galactose NADPH-linked activities also appear to be inducible, since only minor activity with L-arabinose and no activity with D-galactose is detected in extracts of cells grown in D-glucose medium. The NADPH-linked activities obtained with these three sugars may result from the actions of distinctly different enzymes or from a single aldose reductase acting on different substrates. High-performance liquid chromatography and gas-liquid chromatography of in vitro D-xylose, L-arabinose, and D-galactose NADPH-linked reactions confirmed xylitol, L-arabitol, and galactitol as the respective conversion products of these sugars. Unlike xylitol, however, neither L-arabitol nor galactitol would support comparable NAD-linked reaction(s) in cellfree extracts of induced P. tannophilus. Thus, the metabolic pathway of D-xylose diverges from those of L-arabinose or D-galactose following formation of the pentitol.  相似文献   

8.
An NADP(+)-dependent D-xylose dehydrogenase from pig liver cytosol was purified about 2000-fold to apparent homogeneity with a yield of 15% and specific activity of 6 units/mg of protein. An Mr value of 62,000 was obtained by gel filtration. PAGE in the presence of SDS gave an Mr value of 32,000, suggesting that the native enzyme is a dimer of similar or identical subunits. D-Xylose, D-ribose, L-arabinose, 2-deoxy-D-glucose, D-glucose and D-mannose were substrates in the presence of NADP+ but the specificity constant (ratio kcat./Km(app.)) is, by far, much higher for D-xylose than for the other sugars. The enzyme is specific for NADP+; NAD+ is not reduced in the presence of D-xylose or other sugars. Initial-velocity studies for the forward direction with xylose or NADP+ concentrations varied at fixed concentrations of the nucleotide or the sugar respectively revealed a pattern of parallel lines in double-reciprocal plots. Km values for D-xylose and NADP+ were 8.8 mM and 0.99 mM respectively. Dead-end inhibition studies to confirm a ping-pong mechanism showed that NAD+ acted as an uncompetitive inhibitor versus NADP+ (Ki 5.8 mM) and as a competitive inhibitor versus xylose. D-Lyxose was a competitive inhibitor versus xylose and uncompetitive versus NADP+. These results fit better to a sequential compulsory ordered mechanism with NADP+ as the first substrate, but a ping-pong mechanism with xylose as the first substrate has not been ruled out. The presence of D-xylose dehydrogenase suggests that in mammalian liver D-xylose is utilized by a pathway other than the pentose phosphate pathway.  相似文献   

9.
Summary A preliminary test for the enzymatic conversion of D-xylose into xylitol by the intact cells of Candida pelliculosa (xylose reductase) coupled with the intact cells of a formate-utilizing methanogen strain HU (hydrogenase and F420-NADP oxidoreductase) was conducted by using H2 as an electron donor of NADP+. In the system, NADP(H) was well regenerated via the methanogen cells and about 90% conversion of xylose to xylitol (ca. 8 g/l) could be achieved at 35°C and pH 7.5 after 24 h incubation.  相似文献   

10.
Summary The initial steps of D-xylose catabolism inFusarium oxysporum have been studied. The presence of the oxidoreductase pathway for D-xylose catabolism was demonstrated. The enzymes involved, D-xylose reductase and xylitol dehydrogenase, were found to be inducible and relatively specific for D-xylose and xylitol. D-xylose isomerase was not detected.  相似文献   

11.
Wild type Escherichia coli K-12 cannot grow on xylitol and we have been unsuccessful in isolating a mutant directly which had acquired this new growth ability. However, a mutant had been selected previously for growth on L-1,2-propanediol as the sole source of carbon and energy. This mutant constitutively synthesized a propanediol dehydrogenase. Recently, we have found that this dehydrogenase fortuitously converted xylitol to D-xylose which could normally be metabolized by E. coli K-12. In addition, it was also discovered that the D-xylose permease fortuitously transported xylitol into the cell. A second mutant was thus isolated from the L-1,2-propanediol-growing mutant that was constitutive for the enzymes of the D-xylose pathway. This mutant could indeed grow on xylitol as the sole source of carbon and energy, by utilizing the enzymes normally involved in D-xylose and L-1,2-propanediol metabolism.  相似文献   

12.
Summary The production of extracellular xylitol from D-xylose by an efficient xylitol-producing yeast, Candida sp. L-102, was studied in shake flask cultures with different nitrogen sources in the basic salt medium. Maximum xylitol production was obtained with urea as the nitrogen source. A final concentration of 100 g/L of xylitol from 114 g/L D-xylose was obtained from the yeast with an indicated yield of 87.7% (based on D-xylose consumed). The average specific xylitol production rate of 0.46 g/g.h was achieved within 65 hours of incubation using 0.3% urea.  相似文献   

13.
The effect of culture conditions on xylitol production rate was investigated using Candida tropicalis IFO 0618. From the variance analysis of xylitol production rate, it was found that initial yeast extract concentration was highly significant (99%), while the interaction between D-xylose concentration and aeration rate was significant (95%). These results show the importance of initial yeast extract concentration and of the balance between D-xylose concentration and aeration in the production of xylitol. It was also clearly shown that C. tropicalis needed more yeast extract concentration for efficient xylitol production than for its growth. In order to enhance xylitol production rate, culture conditions were optimized by the Box-Wilson method. In this respect, initial D-xylose concentration, yeast extract concentration, and K(L)a were chosen as the independent factors in 2(3)-factorial experimental design. As the result of experiments, a maximum xylitol production rate of 2.67 g/L . h was obtained when initial D-xylose concentration and yeast extract concentration were 172.0 and 21.0 g/L, respectively, and K(L)a was 451.50 h(-1) by 90% oxygen gas. (c) 1992 John Wiley & Sons, Inc.  相似文献   

14.
Pentose fermentation to ethanol with recombinant Saccharomyces cerevisiae is slow and has a low yield. A likely reason for this is that the catabolism of the pentoses D-xylose and L-arabinose through the corresponding fungal pathways creates an imbalance of redox cofactors. The process, although redox neutral, requires NADPH and NAD+, which have to be regenerated in separate processes. NADPH is normally generated through the oxidative part of the pentose phosphate pathway by the action of glucose-6-phosphate dehydrogenase (ZWF1). To facilitate NADPH regeneration, we expressed the recently discovered gene GDP1, which codes for a fungal NADP+-dependent D-glyceraldehyde-3-phosphate dehydrogenase (NADP-GAPDH) (EC 1.2.1.13), in an S. cerevisiae strain with the D-xylose pathway. NADPH regeneration through an NADP-GAPDH is not linked to CO2 production. The resulting strain fermented D-xylose to ethanol with a higher rate and yield than the corresponding strain without GDP1; i.e., the levels of the unwanted side products xylitol and CO2 were lowered. The oxidative part of the pentose phosphate pathway is the main natural path for NADPH regeneration. However, use of this pathway causes wasteful CO2 production and creates a redox imbalance on the path of anaerobic pentose fermentation to ethanol because it does not regenerate NAD+. The deletion of the gene ZWF1 (which codes for glucose-6-phosphate dehydrogenase), in combination with overexpression of GDP1 further stimulated D-xylose fermentation with respect to rate and yield. Through genetic engineering of the redox reactions, the yeast strain was converted from a strain that produced mainly xylitol and CO2 from D-xylose to a strain that produced mainly ethanol under anaerobic conditions.  相似文献   

15.
Wild type Escherichia coli K-12 cannot grow on xylitol and we have been unsuccessful in isolating a mutant directly which had acquired this new growth ability. However, a mutant had been selected previously for growth on L-1,2-propanediol as the sole source of carbon and energy. This mutant constitutively synthesized a propanediol dehydrogenase. Recently, we have found that this dehydrogenase fortuitously converted xylitol to D-xylose which could normally be metabolized by E. coli K-12. In addition, it was also discovered that the D-xylose permease fortuitously transported xylitol into the cell. A second mutant was thus isolated from the L-1,2-propanediol-growing mutant that was constitutive for enzymes of the D-xylose pathway. This mutant could indeed grow on xylitol as the sole source of carbon and energy, by utilizing the enzymes normally involved in D-xylose and L-1,2-propanediol metabolism.  相似文献   

16.
The effect on D-xylose utilization and the corresponding xylitol and ethanol production by Candida shehatae (ATCC 22984) were examined with different nitrogen sources. These included organic (urea, asparagine, and peptone) and inorganic (ammonium chloride, ammonium nitrate, ammonium sulphate, and potassium nitrate) sources. Candida shehatae did not grow on potassium nitrate. Improved ethanol production (Y(p/s), yield coefficient (grams product/grams substrate), 0.34) was observed when organic nitrogen sources were used. Correspondingly, the xylitol production was also higher with organic sources. Ammonium sulphate showed the highest ethanol:xylitol ratio (11.0) among all the nitrogen sources tested. The ratio of NADH- to NADPH-linked D-xylose reductase (EC 1.1.1.21) appeared to be rate limiting during ethanologenesis of D-xylose. The levels of xylitol dehydrogenase (EC 1.1.1.9) were also elevated in the presence of organic nitrogen sources. These results may be useful in the optimization of alcohol production by C. shehatae during continuous fermentation of D-xylose.  相似文献   

17.
Summary Xylitol was produced as a metabolic by-product by a number of yeasts when grown on medium containing D-xylose as carbon and energy sources. Among the yeast strains tested, a mutant strain of Candida tropicalis (HXP2) was found to produce xylitol from D-xylose with a high yield (>90%). Ethanol was also produced by HXP2 when D-glucose, D-fructose, or sucrose were used as substrates. The high-xylitol-producing yeast mutant is a good organism for the production of xylitol from biomass that contains D-xylose.  相似文献   

18.
Seventeen cultures belonging to three genera of facultative bacteria ( Serratia, Cellulomonas, and Corynebacterium) were screened for the production of xylitol, a sugar alcohol used as a sweetener in the pharmaceutical and food industries. The bacterial strains that utilized D-xylose for growth were investigated for xylitol production. A chromogenic assay of both solid and liquid cultures showed that ten of the 17 bacteria screened could grow on D-xylose and produce detectable quantities of xylitol during 24-96 h of fermentation. Among the screened cultures, Corynebacterium sp. B-4247 produced the highest amount of xylitol. In addition, the ten bacterial cultures that initially produced xylitol were studied for the effect of the environmental factors, such as temperature, concentration of D-xylose and aeration, on xylitol production.  相似文献   

19.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

20.
Conversion of pentoses by yeasts   总被引:2,自引:0,他引:2  
The utilization and conversion of D-xylose, D-xylulose, L-arabinose, and xylitol by yeast strains have been investigated with the following results: (1) The majority of yeasts tested utilize D-xylose and produce polyols, ethanol, and organic acids. The type and amount of products formed varies with the yeast strains used. The most commonly detected product is xylitol. (2)The majority of yeasts tested utilize D-xylulose aerobically and fermentatively to produce ethanol, xylitol, D-arabitol, and organic acids. The type and amount of products varies depending upon the yeast strains used. (3) Xylitol is a poor carbon and energy source for most yeasts tested. Some yeast strains produce small amounts of ethanol from xylitol. (4) Most yeast strains utilize L-arabinose, and L-arabitol is the common product. Small amounts of ethanol are also produced by some yeast strains. (5) Of the four substrates examined, D-xylulose was the perferred substrate, followed by D-xylose, L-arabinose, and xylitol. (6) Mutant yeast strains that exhibit different metabolic product patterns can be induced and isolated from Candida sp. Saccharomyces cerevisiae, and other yeasts. These mutant strains can be used for ethanol production from D-xylose as well as for the study of metabolic regulation of pentose utilization in yeasts.  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