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 437 毫秒
1.
Hydroquinone-α-glucoside was synthesised from hydroquinone and maltose as glucosyl donor by transglucosylation in a water system with α-glucosidase from baker’s yeast. Only one phenolic –OH group was α-anomer-selectively glucosylated. The optimum conditions for transglucosylation reaction were at 30 °C for 20 h with 50 mM hydroquinone and 1.5 M maltose in 100 mM sodium citrate/phosphate buffer at pH 5.5. The glucoside was obtained at 0.6 mg/ml with a 4.6% molar yield with respect to hydroquinone.  相似文献   

2.
The ability of an inulosucrase (IS) from Lactobacillus gasseri DSM 20604 to synthesize fructooligosaccharides (FOS) and maltosylfructosides (MFOS) in the presence of sucrose and sucrose-maltose mixtures was investigated after optimization of synthesis conditions, including enzyme concentration, temperature, pH, and reaction time. The maximum formation of FOS, which consist of β-2,1-linked fructose to sucrose, was 45% (in weight with respect to the initial amount of sucrose) and was obtained after 24 h of reaction at 55°C in the presence of sucrose (300 g liter−1) and 1.6 U ml−1 of IS–25 mM sodium acetate buffer–1 mM CaCl2 (pH 5.2). The production of MFOS was also studied as a function of the initial ratios of sucrose to maltose (10:50, 20:40, 30:30, and 40:20, expressed in g 100 ml−1). The highest yield in total MFOS was attained after 24 to 32 h of reaction time and ranged from 13% (10:50 sucrose/maltose) to 52% (30:30 sucrose/maltose) in weight with respect to the initial amount of maltose. Nuclear magnetic resonance (NMR) structural characterization indicated that IS from L. gasseri specifically transferred fructose moieties of sucrose to either C-1 of the reducing end or C-6 of the nonreducing end of maltose. Thus, the trisaccharide erlose [α-d-glucopyranosyl-(1→4)-α-d-glucopyranosyl-(1→2)-β-d-fructofuranoside] was the main synthesized MFOS followed by neo-erlose [β-d-fructofuranosyl-(2→6)-α-d-glucopyranosyl-(1→4)-α-d-glucopyranose]. The formation of MFOS with a higher degree of polymerization was also demonstrated by the transfer of additional fructose residues to C-1 of either the β-2,1-linked fructose or the β-2,6-linked fructose to maltose, revealing the capacity of MFOS to serve as acceptors.  相似文献   

3.
Thermostable Amylolytic Enzymes from a New Clostridium Isolate   总被引:12,自引:9,他引:3       下载免费PDF全文
A new Clostridium strain was isolated on starch at 60°C. Starch, pullulan, maltotriose, and maltose induced the synthesis of α-amylase and pullulanase, while glucose, ribose, fructose, and lactose did not. The formation of the amylolytic enzymes was dependent on growth and occurred predominantly in the exponential phase. The enzymes were largely cell bound during growth of the organism with 0.5% starch, but an increase of the starch concentration in the growth medium was accompanied by the excretion of α-amylase and pullulanase into the culture broth; but also by a decrease of total activity. α-Amylase, pullulanase, and α-glucosidase were active in a broad temperature range (40 to 85°C) and displayed temperature optima for activity at 60 to 70°C. During incubation with starch under aerobic conditions at 75°C for 2 h, the activity of both enzymes decreased to only 90 or 80%. The apparent Km values of α-amylase, pullulanase, and α-glucosidase for their corresponding substrates, starch, pullulan, and maltose were 0.35 mg/ml, 0.63 mg/ml, and 25 mM, respectively.  相似文献   

4.
We studied the prostaglandin (PG) synthetic capacity of microsomes of a relatively pure population of rabbit enterocytes and determined ideal conditions for product synthesis. The epithelial cells were freed from the basement membrane by a combination of calcium chelation and mechanical vibration, and 100,000 x g microsomes were prepared. These microsomes were found to synthesize PG from exogenously added arachidonic acid. The ideal conditions for the reaction were a microsomal protein concentration of 1.0 mg/ml, an arachidonic acid concentration of 33 μM, a reaction mixture pH of 8.0 − 9.5 and with epinephrine 1.5 mM added as a cofactor. The product yields increased linearly with time up to 30 min, of incubation and were inhibited by 100 μM indomethacin. Under the above ideal conditions enterocyte microsomes yielded the following products expressed as pmole/mg protein/20 min, incubation: PGF 98±7, PGE2 48±9, PGD2 28±7, TxB2 40±5, 6 Keto PGF 15 ± 6.  相似文献   

5.
Alpha-glucan phosphorylase catalyzes the reversible cleavage of α-1-4-linked glucose polymers into α-D-glucose-1-phosphate. We report the recombinant production of an α-glucan/maltodextrin phosphorylase (PF1535) from a hyperthermophilic archaeon, Pyrococcus furiosus, and the first detailed biochemical characterization of this enzyme from any archaeal source using a mass-spectrometry-based assay. The apparent 98 kDa recombinant enzyme was active over a broad range of temperatures and pH, with optimal activity at 80 °C and pH 6.5–7. This archaeal protein retained its complete activity after 24 h at 80 °C in Tris-HCl buffer. Unlike other previously reported phosphorylases, the Ni-affinity column purified enzyme showed broad substrate specificity in both the synthesis and degradation of maltooligosaccharides. In the synthetic direction of the enzymatic reaction, the lowest oligosaccharide required for the chain elongation was maltose. In the degradative direction, the archaeal enzyme can produce glucose-1-phosphate from maltotriose or longer maltooligosaccharides including both glycogen and starch. The specific activity of the enzyme at 80 °C in the presence of 10 mM maltoheptaose and at 10 mg ml–1 glycogen concentration was 52 U mg–1 and 31 U mg–1, respectively. The apparent Michaelis constant and maximum velocity for inorganic phosphate were 31 ± 2 mM and 0.60 ± 0.02 mM min–1 µg–1, respectively. An initial velocity study of the enzymatic reaction indicated a sequential bi-bi catalytic mechanism. Unlike the more widely studied mammalian glycogen phosphorylase, the Pyrococcus enzyme is active in the absence of added AMP.  相似文献   

6.
Arbutin (Ab, 4-hydroxyphenyl β-glucopyranoside) is a glycosylated hydroquinone known to prevent the formation of melanin by inhibiting tyrosinase. An arbutin-α-glucoside was synthesized by the transglycosylation reaction of amylosucrase (AS) of Deinococcus geothermalis (DGAS) using arbutin and sucrose as an acceptor and a donor, respectively. The maximum yield of the arbutin transglycosylation product was determined to be over 98% with a 1:0.5 molar ratio of donor and acceptor molecules (sucrose and arbutin), in 50 mM sodium citrate buffer pH 7 at 35 °C. TLC and HPLC analyses revealed that only one transglycosylation product was observed, supporting the result that the transglycosylation reaction of DGAS was very specific. The arbutin transglycosylation product was isolated by preparative recycling HPLC. The structural analyses using 13C and 1H NMR proved that the transglycosylated product was 4-hydroxyphenyl β-maltoside (Ab-α-glucoside), in which a glucose molecule was linked to arbutin via an α-(1 → 4)-glycosidic linkage.  相似文献   

7.

Background

Plant secondary metabolites are possess several biological activities such as anti-mutagenic, anti-carcinogenic, anti-aging, etc. Cell suspension culture is one of the most effective systems to produce secondary metabolites. It is possible to increase the phenolic compounds and tocopherols by using cell suspensions. Studies on tocopherols production by cell suspension cultures are seldom and generally focused on seed oil plants. Although fresh grape, grape seed, pomace and grape seed oil had tocopherols, with our best knowledge, there is no research on tocopherol accumulation in the grape cell suspension cultures. In this study, it was aimed to determine the effects of cadmium chloride treatments on secondary metabolite production in cell suspension cultures of grapevine. Cell suspensions initiated from callus belonging to petiole tissue was used as a plant material. Cadmium chloride was applied to cell suspension cultures in different concentration (1.0 mM and 1.5 mM) to enhance secondary metabolite (total phenolics, total flavanols, total flavonols, trans-resveratrol, and α-, β-, γ- δ-tocopherols) production. Cells were harvested at two days intervals until the 6th day of cultures. Amounts of total phenolics, total flavanols and total flavonols; trans-resveratrol and tocopherols (α-, β-, γ- and δ-tocopherols) and dry cell weights were determined in the harvested cells.

Results

Phenolic contents were significantly affected by the sampling time and cadmium concentrations. The highest values of total phenolic (168.82 mg/100 g), total flavanol (15.94 mg/100 g), total flavonol (14.73 mg/100 g) and trans-resveratrol (490.76 μg/100 g) were found in cells treated with 1.0 mM CdCl2 and harvested at day 2. Contents of tocopherols in the cells cultured in the presence of 1.0 mM CdCl2 gradually increased during the culture period and the highest values of α, β and γ tocopherols (145.61, 25.52 and 18.56 μg/100 g) were detected in the cell cultures collected at day 6.

Conclusions

As a conclusion, secondary metabolite contents were increased by cadmium chloride application and sampling time, while dry cell weights was reduced by cadmium chloride treatments.  相似文献   

8.
Three isoforms of α-glucosidase (EC 3.2.1.20) have been extracted from pea (Pisum sativum L.) seedlings and separated by DEAE-cellulose and CM-Sepharose chromatography. Two α-glucosidase isoforms (αG1 and αG2) were most active under acid conditions, and appeared to be apoplastic. A neutral form (αG3) was most active near pH 7, and was identified as a chloroplastic enzyme. Together, the activity of αG1 and αG2 in apoplastic preparations accounted for 21% of the total acid α-glucosidase activity recovered from pea stems. The vast majority (86%) of the apoplastic acid α-glucosidase activity was due to αG1. The apparent Km values for maltose of αG1 and αG2 were 0.3 and 1.3 millimolar, respectively. The apparent Km for maltose of αG3 was 33 millimolar. The respective native molecular weights of αG1, αG2, and αG3 were 125,000, 150,000, and 110,000.  相似文献   

9.
Pathway for Biodegradation of p-Nitrophenol in a Moraxella sp   总被引:16,自引:10,他引:16       下载免费PDF全文
A Moraxella strain grew on p-nitrophenol with stoichiometric release of nitrite. During induction of the enzymes for growth on p-nitrophenol, traces of hydroquinone accumulated in the medium. In the presence of 2,2′-dipyridyl, p-nitrophenol was converted stoichiometrically to hydroquinone. Particulate enzymes catalyzed the conversion of p-nitrophenol to hydroquinone in the presence of NADPH and oxygen. Soluble enzymes catalyzed the conversion of hydroquinone to γ-hydroxymuconic semialdehyde, which was identified by high-performance liquid chromatography (HPLC)-mass spectroscopy. Upon addition of catalytic amounts of NAD+, γ-hydroxymuconic semialdehyde was converted to β-ketoadipic acid. In the presence of pyruvate and lactic dehydrogenase, substrate amounts of NAD were required and γ-hydroxymuconic semialdehyde was converted to maleylacetic acid, which was identified by HPLC-mass spectroscopy. Similar results were obtained when the reaction was carried out in the presence of potassium ferricyanide. Extracts prepared from p-nitrophenol-growth cells also contained an enzyme that catalyzed the oxidation of 1,2,4-benzenetriol to maleylacetic acid. The enzyme responsible for the oxidation of 1,2,4-benzenetriol was separated from the enzyme responsible for hydroquinone oxidation by DEAE-cellulose chromatography. The results indicate that the pathway for biodegradation of p-nitrophenol involves the initial removal of the nitro group as nitrite and formation of hydroquinone. 1,4-Benzoquinone, a likely intermediate in the initial reaction, was not detected. Hydroquinone is converted to β-ketoadipic acid via γ-hydroxymuconic semialdehyde and maleylacetic acid.  相似文献   

10.
α-Glucosidase from Bacillus stearothermophilus was used as a catalyst for oligosaccharide synthesis by reversed hydrolysis. The yield of disaccharides and trisaccharides depended strongly on the units of enzyme activity added, and on the stability of the enzyme under reaction conditions. When glucose was the only saccharide present in the reaction mixture with α-glucosidase, isomaltose (51%), nigerose (25%), maltose (14%) and kojibiose (10%) were formed. In 50% glucose solution, disaccharide concentrations reached up to 400 mmol/l and trisaccharides were also produced. When other saccharides (mannose or xylose), in addition to glucose, were present in the reaction mixture, both homodisaccharides and heterodisaccharides were formed, their quantity being dependent on the glucose/saccharide acceptor ratios. The highest yields of oligosaccharides were observed with glucose alone, consistent with the observation that the enzyme stability was highest with glucose as the sole saccharide.  相似文献   

11.
A β-phosphoglucomutase (β-PGM) mutant of Lactococcus lactis subsp. lactis ATCC 19435 was constructed using a minimal integration vector and double-crossover recombination. The mutant and the wild-type strain were grown under controlled conditions with different sugars to elucidate the role of β-PGM in carbohydrate catabolism and anabolism. The mutation did not significantly affect growth, product formation, or cell composition when glucose or lactose was used as the carbon source. With maltose or trehalose as the carbon source the wild-type strain had a maximum specific growth rate of 0.5 h−1, while the deletion of β-PGM resulted in a maximum specific growth rate of 0.05 h−1 on maltose and no growth at all on trehalose. Growth of the mutant strain on maltose resulted in smaller amounts of lactate but more formate, acetate, and ethanol, and approximately 1/10 of the maltose was found as β-glucose 1-phosphate in the medium. Furthermore, the β-PGM mutant cells grown on maltose were considerably larger and accumulated polysaccharides which consisted of α-1,4-bound glucose units. When the cells were grown at a low dilution rate in a glucose and maltose mixture, the wild-type strain exhibited a higher carbohydrate content than when grown at higher growth rates, but still this content was lower than that in the β-PGM mutant. In addition, significant differences in the initial metabolism of maltose and trehalose were found, and cell extracts did not digest free trehalose but only trehalose 6-phosphate, which yielded β-glucose 1-phosphate and glucose 6-phosphate. This demonstrates the presence of a novel enzymatic pathway for trehalose different from that of maltose metabolism in L. lactis.  相似文献   

12.
A neutral α-glucosidase (EC 3.2.1.20) activity was shown to be associated with granules which are sedimentable at 10 000 g after differential centrifugation of mouse peritoneal macrophage homogenates. When the post-nuclear supernatant was centrifuged in a sucrose density gradient, high activities for neutral α-glucosidase and β-glucuronidase (EC 3.2.1.31) were detected in the bottom fractions because of aggregation of the granules. Neutral α-glucosidase-containing granules were completely disaggregated by the addition of 20 units/ml of heparin and 10 mM Tris-HCl (pH 7.2), which caused only a partial disaggregation of β-glucuronidase-containing granules. The addition of a high concentration of heparin, Tris buffer, or KCl to the gradient gave the same patterns of disaggregation of the granules. Under the condition in which about 50% of the total β-glucuronidase activity was released into the medium, depending on phagocytosis, very little α-glucosidase was released. These observations suggested that neutral α-glucosidase may localize in non-lysosomal granules.  相似文献   

13.
The effects of operating conditions on the enzymatic hydrolysis of corn starch were investigated. A commercial α-amylase produced by Bacillus sp. was used for the hydrolysis experiments. The degree of starch hydrolysis (%) and residual α-amylase activity (%) was investigated versus process variables, including pH, temperature, viscosity, impeller speed, processing time and some materials added such as hydrolysate, maltose, glucose, ethanol and CaCl2 using a stirred batch reactor. The mathematical models depending on the operating conditions were also derived using the experimental data of residual starch concentration. Some inactivation models were tested to determine the relationship between process variables and enzyme stability during the hydrolysis process.  相似文献   

14.
Kinetic mechanism and stoichiometry of scavenging the 2,2-diphenyl-1-picrylhydrazyl radical by glutathione and its novel analog, containing α-glutamyl residue in place of the γ-glutamyl moiety, were studied using different ratios of reagents. At low concentrations of the peptides, the process was described as a bimolecular reaction obeying the stoichiometric ratio 1:1. However, at excess of peptides the formation of a non-covalent complex between the reagents was discovered and characterized by dissociation constants K = 0.61 mM for glutathione and K = 0.27 mM for the glutathione α-glutamyl analog, respectively. The complex formation was followed by a reaction step that was characterized by the similar rate constant k = 0.02 s−1 for both peptides. Thus, the apparently different antioxidant activity of these two peptides, observed under common assay conditions, was determined by differences in the formation of this non-covalent complex.  相似文献   

15.

Background

Arabidopsis plants accumulate maltose from starch breakdown during cold acclimation. The Arabidopsis mutant, maltose excess1-1, accumulates large amounts of maltose in the plastid even in the warm, due to a deficient plastid envelope maltose transporter. We therefore investigated whether the elevated maltose level in mex1-1 in the warm could result in changes in metabolism and physiology typical of WT plants grown in the cold.

Principal Findings

Grown at 21 °C, mex1-1 plants were much smaller, with fewer leaves, and elevated carbohydrates and amino acids compared to WT. However, after transfer to 4 °C the total soluble sugar pool and amino acid concentration was in equal abundance in both genotypes, although the most abundant sugar in mex1-1 was still maltose whereas sucrose was in greatest abundance in WT. The chlorophyll a/b ratio in WT was much lower in the cold than in the warm, but in mex1-1 it was low in both warm and cold. After prolonged growth at 4 °C, the shoot biomass, rosette diameter and number of leaves at bolting were similar in mex1-1 and WT.

Conclusions

The mex1-1 mutation in warm-grown plants confers aspects of cold acclimation, including elevated levels of sugars and amino acids and low chlorophyll a/b ratio. This may in turn compromise growth of mex1-1 in the warm relative to WT. We suggest that elevated maltose in the plastid could be responsible for key aspects of cold acclimation.  相似文献   

16.
A bacterial glucoamylase was purified from the anaerobic thermophilic bacterium Clostridium thermosaccharolyticum and characterized. The enzyme, which was purified 63-fold, with a yield of 36%, consisted of a single subunit with an apparent molecular mass of 75 kDa. The purified enzyme was able to attack α-1,4- and α-1,6-glycosidic linkages in various α-glucans, liberating glucose with a β-anomeric configuration. The purified glucoamylase, which was optimally active at 70°C and pH 5.0, attacked preferentially polysaccharides such as starch, glycogen, amylopectin, and maltodextrin. The velocity of oligosaccharide hydrolysis decreased with a decrease in the size of the substrate. The Km values for starch and maltose were 18 mg/ml and 20 mM, respectively. Enzyme activity was not significantly influenced by Ca2+, EDTA, or α- or β-cyclodextrins.  相似文献   

17.
The kinetics for the reduction of sulfate alone and for concurrent uranium [U(VI)] and sulfate reduction, by mixed and pure cultures of sulfate-reducing bacteria (SRB) at 21 ± 3°C were studied. The mixed culture contained the SRB Desulfovibrio vulgaris along with a Clostridium sp. determined via 16S ribosomal DNA analysis. The pure culture was Desulfovibrio desulfuricans (ATCC 7757). A zero-order model best fit the data for the reduction of sulfate from 0.1 to 10 mM. A lag time occurred below cell concentrations of 0.1 mg (dry weight) of cells/ml. For the mixed culture, average values for the maximum specific reaction rate, Vmax, ranged from 2.4 ± 0.2 μmol of sulfate/mg (dry weight) of SRB · h−1) at 0.25 mM sulfate to 5.0 ± 1.1 μmol of sulfate/mg (dry weight) of SRB · h−1 at 10 mM sulfate (average cell concentration, 0.52 mg [dry weight]/ml). For the pure culture, Vmax was 1.6 ± 0.2 μmol of sulfate/mg (dry weight) of SRB · h−1 at 1 mM sulfate (0.29 mg [dry weight] of cells/ml). When both electron acceptors were present, sulfate reduction remained zero order for both cultures, while uranium reduction was first order, with rate constants of 0.071 ± 0.003 mg (dry weight) of cells/ml · min−1 for the mixed culture and 0.137 ± 0.016 mg (dry weight) of cells/ml · min−1 (U0 = 1 mM) for the D. desulfuricans culture. Both cultures exhibited a faster rate of uranium reduction in the presence of sulfate and no lag time until the onset of U reduction in contrast to U alone. This kinetics information can be used to design an SRB-dominated biotreatment scheme for the removal of U(VI) from an aqueous source.  相似文献   

18.
A novel enzyme, β-phenylalanine ester hydrolase, useful for chiral resolution of β-phenylalanine and for its β-peptide synthesis was characterized. The enzyme purified from the cell free-extract of Sphingobacterium sp. 238C5 well hydrolyzed β-phenylalanine esters (S)-stereospecifically. Besides β-phenylalanine esters, the enzyme catalyzed the hydrolysis of several α-amino acid esters with l-stereospecificity, while the deduced 369 amino acid sequence of the enzyme exhibited homology to alkaline d-stereospecific peptide hydrolases from Bacillus strains. Escherichia coli transformant expressing the β-phenylalanine ester hydrolase gene exhibited an about 8-fold increase in specific (S)-β-phenylalanine ethyl ester hydrolysis as compared with that of Sphingobacterium sp. 238C5. The E. coli transformant showed (S)-enantiomer specific esterase activity in the reaction with a low concentration (30 mM) of β-phenylalanine ethyl ester, while it showed both esterase and transpeptidase activity in the reaction with a high concentration (170 mM) of β-phenylalanine ethyl ester and produced β-phenylalanyl-β-phenylalanine ethyl ester. This transpeptidase activity was useful for β-phenylalanine β-peptide synthesis.  相似文献   

19.
An alternative and fast method for the purification of an exo-β- -galactofuranosidase has been developed using a 4-aminophenyl 1-thio-β- -galactofuranoside affinity chromatography system and specific elution with 10 mM -galactono-1,4-lactone in a salt gradient. A concentrated culture medium from Penicillium fellutanum was chromatographed on DEAE–Sepharose CL 6B followed by chromatography on the affinity column, yielding two separate peaks of enzyme activity when elution was performed with 10 mM -galactono-1,4-lactone in a 100–500 mM NaCl salt gradient. Both peaks behaved as a single 70 kDa protein, as detected by SDS-PAGE. Antibodies elicited against a mixture of the single bands excised from the gel were capable of immunoprecipitating 0.2 units out of 0.26 total units of the enzyme from a crude extract. The glycoprotein nature of the exo-β- -galactofuranosidase was ascertained through binding to Concanavalin A–Sepharose as well as by specific reaction with Schiff reagent in Western blots. The purified enzyme has an optimum acidic pH (between 3 and 6), and Km and Vmax values of 0.311 mM and 17 μmol h−1 μg−1 respectively, when 4-nitrophenyl β- -galactofuranoside was employed as the substrate.  相似文献   

20.
Previous reports have suggested the primary mode of action of the allelochemical hydroquinone involves disruption of root cell membrane transport. Here we report the effects of hydroquinone on common bean (Phaseolus vulgaris) plants. Growth of leaves, roots and stems were all inhibited by 14 day exposure to 0.01 mM or 0.25 mM hydroquinone. Chlorophyll fluorescence (Fv/Fm) was inhibited by 0.25 mM hydroquinone. The membrane potential of P. vulgaris root cortex cells briefly hyperpolarized and subsequently slowly transiently depolarized upon abrupt exposure to a range of hydroquinone concentrations. Both the hyperpolarization and depolarization were concentration dependent but appeared saturable. Root cells exposed to 0.03 mM hydroquinone hyperpolarized 3.4 mV (+/− 0.6 s.e.) 3 minutes after the start of exposure then depolarized 36.7 mV (+/− 3.9) with no effect evident after 24 hours. Individual recordings showed a response to as little as 0.001 mM hydroquinone. Exposure of P. vulgaris root cells to arbutin, a nontoxic monoglucoside of hydroquinone, produced a similar but much smaller (approximately 25%) electrical response. Exposure of root cells of Antennaria microphylla, a known allelopathic source (donor plant) of hydroquinone, also produced a much smaller hyperpolarization and depolarization response. It is concluded that the electrical response to hydroquinone by P. vulgaris root cells and the changes in membrane transport they represent are not sufficiently large or long lasting enough to disrupt mineral and water uptake leading to plant injury. The possibility, however, that these events are related to initiation of signal transduction events leading to cell death is discussed.Key words: allelopathy, hydroquinone, membrane potential, depolarization, hyperpolarization, Phaseolus vulgaris, Antennaria microphylla  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