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 46 毫秒
1.
Xylose reductase (XR) is a key enzyme in D-xylose metabolism, catalyzing the reduction of D-xylose to xylitol. An NADH-preferring XR was purified to homogeneity from Candida parapsilosis KFCC-10875, and the xyl1 gene encoding a 324-amino-acid polypeptide with a molecular mass of 36,629 Da was subsequently isolated using internal amino acid sequences and 5' and 3' rapid amplification of cDNA ends. The C. parapsilosis XR showed high catalytic efficiency (kcat/Km = 1.46 s(-1) mM(-1)) for D-xylose and showed unusual coenzyme specificity, with greater catalytic efficiency with NADH (kcat/Km = 1.39 x 10(4) s(-1) mM(-1)) than with NADPH (kcat/Km = 1.27 x 10(2) s(-1) mM(-1)), unlike all other aldose reductases characterized. Studies of initial velocity and product inhibition suggest that the reaction proceeds via a sequentially ordered Bi Bi mechanism, which is typical of XRs. Candida tropicalis KFCC-10960 has been reported to have the highest xylitol production yield and rate. It has been suggested, however, that NADPH-dependent XRs, including the XR of C. tropicalis, are limited by the coenzyme availability and thus limit the production of xylitol. The C. parapsilosis xyl1 gene was placed under the control of an alcohol dehydrogenase promoter and integrated into the genome of C. tropicalis. The resulting recombinant yeast, C. tropicalis BN-1, showed higher yield and productivity (by 5 and 25%, respectively) than the wild strain and lower production of by-products, thus facilitating the purification process. The XRs partially purified from C. tropicalis BN-1 exhibited dual coenzyme specificity for both NADH and NADPH, indicating the functional expression of the C. parapsilosis xyl1 gene in C. tropicalis BN-1. This is the first report of the cloning of an xyl1 gene encoding an NADH-preferring XR and its functional expression in C. tropicalis, a yeast currently used for industrial production of xylitol.  相似文献   

2.
Two 2,5-diketo-d-gluconate reductases, I and II, were purified respectively 918-fold and 28-fold from a mutant strain derived from Corynebacterium sp. SHS 0007. The enzymes appeared to be homogeneous on polyacrylamide gel electrophoresis. Both reductases converted 2,5-diketo-d-gluconate to 2-keto-l-gulonate in the presence of NADPH and seemed to be active only for reduction. The molecular weights of reductases I and II were estimated to be 29,000 and 34,000, respectively; and both were monomeric. Their isoelectric points were respectively pH 4.3 and pH 4.1. The optimum pH was 6.0 to 7.0 for reductase I, and 6.0 to 7.5 for reductase II. The Km values (pH 7.0, 30°C) of reductase I for 2,5-diketo-d-gluconate and for NADPH were 1.8 mM and 12 μM, respectively; and the corresponding values of reductase II were 13.5 mM and 13 μM. Both reductases converted 5-keto-d-fructose to l-sorbose in the presence of NADPH.  相似文献   

3.
NADPH-cytochrome P-450 reductases from pig liver and kidney and rabbit liver microsomes were purified to a specific activity of 50–62 μmol cytochrome c reduced/min/mg. All reductase preparations were separated into one major and one minor fraction on Sephadex G-200 columns. The molecular weights of the major fractions of the reductases were estimated to be 74,000, 75,000, and 75,500 for rabbit liver, pig kidney, and liver reductases, respectively, whereas the molecular weight of the minor fractions of these reductases, 67,000, was the same as that of the steapsin-solubilized pig liver reductase on SDS-polyacrylamide gel electrophoresis. Km values for NADPH and cytochrome c were: 20 and 29 μm or 14 and 28 μm for the pig kidney or liver reductase, respectively. Immunochemical studies, including Ouchterlony double diffusion experiments and inhibition of benzphetamine N-demethylation activity in microsomes by antibody against pig liver NADPH-cytochrome P-450 reductase, indicated the similarity of the purified liver and kidney reductases. There were no differences in the ability to reconstitute NADPH-mediated benzphetamine N-demethylation and laurate hydroxylation in reconstituted systems between the pig liver and kidney reductases, indicating that the reductase did not determine substrate specificity in these systems.  相似文献   

4.
A yeast growing at 48°C was isolated from soil and the strain was identified as Cryptococcus lactativorus. The aldose reductase which the strain produced was purified 114-fold with an overall recovery of 36%. The stability of the enzyme was higher than that of other aldose reductases. The half life of the enzyme was 800 h and 14 h at 30°C and 50°C, respectively. The enzyme showed the best activity with d-xylose. l-Sorbose and d-fructose were also reduced by the enzyme. The enzyme was active with both NADPH and NADH as a conenzyme, and the activity with NADH was 1.25 times higher than that with NADPH. The Kmapp value for d-xylose was 8.6 mM and the Vmaxapp was 20.8 units/mg NADH was used as a coenzyme. The Kmapp values for NADPH and NADH were 6μM and 170 μM, respectively, when d-glucose was used as a substrate.  相似文献   

5.
A xylose reductase (XR) gene was identified from the Neurospora crassa whole-genome sequence, expressed heterologously in Escherichia coli, and purified as a His6-tagged fusion in high yield. This enzyme is one of the most active XRs thus far characterized and may be used for the in vitro production of xylitol.  相似文献   

6.
Xylose reductase (XR) is a key enzyme in xylose metabolism because it catalyzes the reduction of xylose to xylitol. In order to study the characteristics of XR from Candida tropicalis SCTCC 300249, its XR gene (xyll) was cloned and expressed in Escherichia coli BL21 (DE3). The fusion protein was purified effectively by Ni2+-chelating chromatography, and the kinetics of the recombinant XR was investigated. The Km values of the C. tropicalis XR for NADPH and NADH were 45.5 μM and 161.9 μM, respectively, which demonstrated that this XR had dual coenzyme specificity. Moreover, this XR showed the highest catalytic efficiency (kcat=1.44×l04 min−1) for xylose among the characterized aldose reductases. Batch fermentation was performed with Saccharomyces serivisiae W303-lA:pYES2XR, and resulted in 7.63 g/L cell mass, 93.67 g/L xylitol, and 2.34 g/L · h xylitol productivity. This XR coupled with its dual coenzyme specificity, high activity, and catalytic efficiency proved its utility in in vitro xylitol production.  相似文献   

7.
Xylose reductase (XR) is a key enzyme in biological xylitol production, and most XRs have broad substrate specificities. During xylitol production from biomass hydrolysate, non-specific XRs can reduce l-arabinose, which is the second-most abundant hemicellulosic sugar, to the undesirable byproduct arabitol, which interferes with xylitol crystallization in downstream processing. To minimize the flux from l-arabinose to arabitol, the l-arabinose-preferring, endogenous XR was replaced by a d-xylose-preferring heterologous XR in Candida tropicalis. Then, Bacillus licheniformis araA and Escherichia coli araB and araD were codon-optimized and expressed functionally in C. tropicalis for the efficient assimilation of l-arabinose. During xylitol fermentation, the control strains BSXDH-3 and KNV converted 9.9 g l-arabinose l−1 into 9.5 and 8.3 g arabitol l−1, respectively, whereas the recombinant strain JY consumed 10.5 g l-arabinose l−1 for cell growth without forming arabitol. Moreover, JY produced xylitol with 42 and 16% higher productivity than BSXDH-3 and KNV, respectively.  相似文献   

8.
Caragana korshinskii hemicelluloses were isolated with 10% KOH at 25 °C for 10 h from the delignified materials. The alkali-extractable hemicelluloses were then successively sub-fractionated by graded precipitation at final ethanol concentrations of 10%, 20%, 30%, 45%, 60%, and 80%, respectively. Neutral sugars and molecular weight analyses of the six hemicellulosic subfractions revealed that the molecular weights and the distribution of branches along the xylan backbone are different among the hemicellulosic fractions obtained in various ethanol concentrations. The less branched hemicelluloses with large molecules were precipitated in lower ethanol percentages, while with the increasing ethanol concentrations, more branched hemicelluloses with low molecular weights were obtained. 1H and 13C NMR studies revealed that the hemicellulosic subfraction precipitated at an ethanol concentration of 45% had a backbone of d-xylose residues and were branched mainly through 4-O-methyl-α-d-glucopyranosyl units.  相似文献   

9.
Proteoglycans from bovine tracheal cartilage were digested with trypsin and chymotrypsin by procedures similar to those described by Mathews (Biochem. J.125, 37 (1971)). Chondroitin sulfate-peptide fragments in the digest were precipitated with cetylpyridinium chloride and subsequently fractionated on a preparative Sepharose 6B column. The fragments, which emerged from the column as a broad peak, were divided into five fractions. Rechromatography of these fractions on an analytical Sepharose 6B column indicated that they had Kav values from 0.17 (fraction 1) to 0.62 (fraction 5). The weight average molecular weight values obtained by meniscus depletion equilibrium centrifugation were 193,000, 126,000, 80,000, 46,000, and 23,000 for fractions 1 to 5, respectively. Values for the molecular weights and for the limiting viscosity numbers, [η], of the fractions were used to determine estimates for α of 0.40–0.46 and for K of 0.43–0.88 in the equation [η] = K·Mvα. These values for α are consistent with a branched structure for the chondroitin sulfate fractions. Papain digests of each of the fractions were chromatographed on Sephadex G-200. The observed distributions of the monomer chains released by this protease were almost the same for each sample, which indicates that the individual chondroitin sulfate chains in all of the original fractions had nearly the same average molecular weights. The data in sum indicate that peptide fragments which contain from 1 to 8 polysaccharide chains are released when the proteoglycans are digested with trypsin-chymotrypsin.Analytical data indicated that all fractions contained 3–11% of their polysaccharide as keratan sulfate. This indicates either that about 50% of the keratan sulfate chains in the original proteoglycan molecules are located in close proximity to the chondroitin sulfate chains or that some peptides contain large numbers of keratan sulfate chains. Proteoglycan preparations which differed by a factor of about 6 in their ratio of chondroitin sulfate to protein yielded very similar elution patterns on Sepharose 6B after trypsin-chymotrypsin digestion.  相似文献   

10.
The kinetics of D-xylose transport were studied in Rhodotorula glutinis. Analysis of the saturation isotherm revealed the presence of at least two carriers for d-xylose in the Rhodotorula plasma membrane. These two carriers exhibited Km values differing by more than an order of magnitude. The low Km carrier was repressed in rapidly growing cells and depressed by starvation of the cells.Several hexoses were observed to inhibit d-xylose transport. In the studies reported here, the inhibitions produced by d-galactose and 2-deoxy-d-glucose were examined in some detail in order to define the interactions of these sugars with the d-xylose carriers. 2-Deoxy-d-glucose competitively inhibited both of the d-xylose carriers. In contrast, only the low-Km carrier was competitively inhibited by d-galactose.  相似文献   

11.
The lignin-carbohydrate complex (LCC-W), isolated from the milled-wood, lignin fraction of Pinus densiflora Sieb. et Zucc., comprised three fractions (W-1,2,3) by gel filtration on Sepharose 4B. W-1 was eluted at the void volume, whereas W-2 and W-3 were included in the gel and had apparent weight-average molecular weights of 5.0 × 105 and 5.0 × 103, respectively. W-2 and W-3 were homogeneous in ultracentrifugal and electrophoretic analyses. The sedimentation coefficients of W-2 and W-3 were 25.7 and 0.4S, respectively. The chemical composition of W-2 was 38.0% of neutral sugar, 6.2% of uronic acid, 51.5% of lignin, and the corresponding values for W-3 were 73.1, 11.0, and 22.2%. The neutral carbohydrate residues of W-2 and W-3 were l-arabinose, d-xylose, d-mannose, d-galactose, and d-glucose in the ratios 15.8:16.2:37.3:16.7:14.0 and 27.6:16.5:26.1:19.3:10.5, respectively. Based on the results of methylation and Smith-degradation analyses, the carbohydrate moiety of the LCC-W fractions was found to be multiply branched. The major backbone structure was composed of (1→4)-linked d-mannopyranosyl residues. By hydrophobicinteraction chromatography on Phenyl- and Octyl-Sepharsoe CL-4B gels, it is concluded that the LCC-W fractions have a hydrophobic property that is exclusively ascribed to the lignin moiety.  相似文献   

12.
The linear arrangement of the three fragments of Ca2+-ATPase from rabbit skeletal muscle sarcoplasmic reticulum with molecular weights of 20,000, 30,000, and 45,000 obtained by limited tryptic hydrolysis was determined by locating the NH2-terminal acetylated methionyl residue of the original peptide in the Mr = 20,000 fragment. Since both the Mr = 20,000 and 30,000 polypeptides originate from a Mr = 55,000 fragment which is distinct from the Mr = 45,000 polypeptide, the sequence of these three fragments was determined to be 20,000, 30,000, and 45,000. The Mr = 20,000 fragment was further cleaved by cyanogen bromide to yield a Mr = 7,000 COOH-terminal fragment which is relatively hydrophilic. The NH2-terminal portion is rich in glutamyl residues. The COOH-terminus of the Mr = 30,000 fragment was determined by both digestion with carboxypeptidases and cyanogen bromide cleavage. Using the partial amino acid sequence of the Ca2+-ATPase, it was deduced that the active site phosphoaspartyl residue is 154 amino acids from the COOH-terminus of the Mr = 30,000 fragment and hence approximately 35,000 Mr from the NH2-terminus of the original Ca2+-ATPase molecule. Furthermore, it was shown that the two tryptic cleavages of the Ca2+-ATPase generating these three large fragments were both single hydrolyses of arginylalanine peptide bonds.  相似文献   

13.
Two ferredoxin-NADP+ reductases (FNRs I and II) [EC 1.6.7.1] were purified from a blue-green alga, Spirulina platensis, by (NH4)2SO4 fractionation, gel filtration on Sephadex G-100 and DEAE-Sephadex A-50 chromatography. FNRs I and II were both FAD-containing enzymes with molecular weights of 33,000, and could photochemically reduce NADP+ to the same extent in the presence of S. platensis ferredoxin, using FNR-depleted membrane fragments of S. platensis. They had similar physical and enzymatic properties, except for chemical properties such as the amino (N)-terminal sequences and the patterns of their peptide maps. The significance of the presence of two FNRs in S. platensis as as of the multiple forms found in other organisms is discussed.  相似文献   

14.
DNA fragments of several sizes have been produced by shearing E. coli DNA under different pressures. These fragments have been used to demonstrate that column chromatography on agarose Bio-Gel A-15M can provide a rapid, inexpensive fractionation and sizing method for single-stranded nucleic acids having masses between 105 and 106 daltons. Both chromatographic and electrophoretic analysis of the sheared DNA indicated that discrete fragment populations were produced at each shearing pressure and that these fragments were distributed essentially symmetrically around a mean piece size. The average molecular weight of the several DNA fragment distributions was determined electrophoretically by comparison with standard DNA fragments obtained from restriction endonuclease cleavage of SV40 viral DNA. The molecular weights of the denatured, sheared fragments (single-stranded) ranged from 1.25 × 105 to 7.4 × 105. The single-stranded DNA fragments were chromatographed over agarose Bio-Gel A-15M and a linear relationship was found to exist between the mobilities and logarithms of the molecular weights. Readily available tRNA, 5s RNA, and φX174 single-stranded circular DNA chromatographed at the extremes of the linear relationship and could be used to calibrate the column chromatography.  相似文献   

15.
Coccidiosis in chickens causes intestinal mucosal lesions and disrupts its integrity leading to a disturbance in absorption of dietary components. The d-xylose absorption test is a sensitive tool of measuring the absorption capacity of the intestine in diseased chickens. In an experiment on broilers, the influence of different anticoccidials on the intestinal absorption capacity of the birds challenged with experimental coccidiosis was evaluated, using the d-xylose absorption test. The experiment had 5 groups of 10 Ross male broiler chickens (24-days-old) as follows: Group 1— negative control received no Eimeria oocystes, Group 2—positive control challenged with mixed Eimeria oocystes, Group 3—positive control dosed with an attenuated oral coccidiosis vaccine, Group 4—positive control dosed with 25 ppm toltrazuril in drinking water and Group 5—positive control received 66 ppm salinomycin sodium, in the diet. The d-xylose absorption test was carried out 5 days after the coccidial infection. Results showed that coccidiosis highly reduced the plasma d-xylose peak level of Group 2 when compared with Group 1 (31 mg/dl at 90 min versus 50 mg/dl at 30–60 min after the d-xylose administration, respectively). The concentration of d-xylose followed cubic (P<0.001, r2 = 0.886) and quadratic (P<0.001, r2 = 0.686) correlations with time in Group 1 and 2, respectively. Anticoccidials enhanced the uptake of d-xylose in the infected birds. The plasma d-xylose reached to its peak in Group 3, 4, and 5 (38.9, 50 and 47.0 mg/dl, respectively) at 60–90 min after the d-xylose administration and had quadratic functions with time (r2 = 0.802, 0.883 and 0.860, respectively, P<0.001). The d-xylose absorption test was a sensitive test for evaluating the influence of anticoccidials on the absorption capacity of intestinal mucosae during coccidiosis in broiler chickens.  相似文献   

16.
A trypsin was purified from the hepatopancreas of snakehead (Channa argus) by ammonium sulfate fractionation and a series of column chromatographies including DEAE-Sepharose, Sephacryl S-200 HR and Hi-Trap Capto-Q. The molecular mass of the purified trypsin was about 22 kDa, as estimated by SDS-PAGE. The optimum pH and temperature of the purified trypsin were 9.0 and 40 °C, respectively. The trypsin was stable in the pH range of 7.5-9.5 and below 45 °C. The enzymatic activity was strongly inhibited by serine proteinase inhibitors, such as MBTI, Pefabloc SC, PMSF, LBTI and benzamidine. Peptide mass fingerprinting (PMF) of the purified protein obtained 2 peptide fragments with 25 amino acid residues and were 100% identical to the trypsinogen from pufferfish (Takifugu rubripes). The activation energy (Ea) of this enzyme was 24.65 kJ·M− 1. Apparent Km was 1.02 μM and kcat was 148 S− 1 for fluorogenic substrate Boc-Phe-Ser-Arg-MCA. A trypsinogen gene encoding 247 amino acid residues was further cloned on the basis of the sequence obtained from PMF and the conserved site peptide of trypsinogen together with 5′-RACE and 3′-RACE. The deduced amino acid sequence contains a signal peptide of 15 residues and an activation peptide of 9 amino acid residues with a mature protein of 223 residues. The catalytic triad His-64, Asp-107, Ser-201 and 12 Cys residues which may form 6 disulfide bonds were conserved. Compared with the PMF data, only 2 amino acid residues difference were identified, suggesting the cloned trypsinogen is quite possibly the precursor of the purified trypsin.  相似文献   

17.
《Phytochemistry》1986,25(11):2463-2469
NADH-nitrite and -hydroxylamine reductases were co-purified from Derxia gummosa. The stoichiometries for the reduction of nitrite and hydroxylamine to ammonia were 3 NADH:1 NO2:1 NH3 and 1 NADH:1 NO2:1 NH3. The Km values for nitrite and hydroxylamine were 4.8 μM and 5.3 mM, respectively, and for NADH they were 6.3 μM for nitrite reductase and 150 μM for hydroxylamine reductase. The optimal pH value for both enzyme activities was 8.5. Both activities were inhibited by NADH in the absence of the appropriate substrate, namely nitrite or hydroxylamine. Studies with amino acid modifiers indicate that histidine, glutamate/aspartate, sulphydryl and tyrosine are essential components of the enzyme protein. Kinetic studies show that nitrite and hydroxylamine were competitive for the same binding site on the enzyme. The results indicate that although nitrite and hydroxylamine reductases are associated with the same enzyme, its main function is the reduction of nitrite to ammonia. Azaserine inhibited the induction of the enzyme.  相似文献   

18.
《Carbohydrate research》1986,148(2):321-330
Two endo-(1→4)-β-d-xylanases (xylanases 1 and 2), which were constitutively synthesised by the fungus Trichoderma koningii, were purified to homogeneity on gel-filtration media and by isoelectric focusing. They had molecular weights of 29,000 (xylanase 1) and 18,000 (xylanase 2), and isoelectric pHs of 7.24 (xylanase 1) and 7.3 (xylanase 2); neither enzyme was associated with carbohydrate. Xylanase 1 had an optimum at the remarkably high temperature of 60–65°. Each enzyme liberated a different range of oligosaccharides from oat-straw arabinoxylan, but only xylanase 1 released l-arabinose and d-xylose. Both xylanases were free from cellulase activity.  相似文献   

19.
Aspergillus carbonarius accumulates xylitol when it grows on d-xylose. In fungi, d-xylose is reduced to xylitol by the NAD(P)H-dependent xylose reductase (XR). Xylitol is then further oxidized by the NAD+-dependent xylitol dehydrogenase (XDH). The cofactor impairment between the XR and XDH can lead to the accumulation of xylitol under oxygen-limiting conditions. Most of the XRs are NADPH dependent and contain a conserved Ile-Pro-Lys-Ser motif. The only known naturally occurring NADH-dependent XR (from Candida parapsilosis) carries an arginine residue instead of the lysine in this motif. In order to overcome xylitol accumulation in A. carbonarius a Lys-274 to Arg point mutation was introduced into the XR with the aim of changing the specificity toward NADH. The effect of the genetic engineering was examined in fermentation for citric acid production and xylitol accumulation by using d-xylose as the sole carbon source. Fermentation with the mutant strain showed a 2.8-fold reduction in xylitol accumulation and 4.5-fold increase in citric acid production compared to the wild-type strain. The fact that the mutant strain shows decreased xylitol levels is assumed to be associated with the capability of the mutated XR to use the NADH generated by the XDH, thus preventing the inhibition of XDH by the high levels of NADH and ensuring the flux of xylose through the pathway. This work shows that enhanced production of citric acid can be achieved using xylose as the sole carbon source by reducing accumulation of other by-products, such as xylitol.  相似文献   

20.
C-and N-polyhedrins from a cytoplasmic polyhedrosis virus (a double-stranded RNA virus) and a nuclear polyhedrosis virus (a DNA virus), respectively, of Orgyia pseudotsugata were compared. Although both polyhedrins appear to stabilize their respective virions and have similar molecular weights, they differed in amino acid composition, tryptic peptide elution profiles from a cation-exchange resin, and N-terminal amino acid sequence and showed no antigenic relatedness. This suggests that these two proteins originated independently of one another.  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