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 31 毫秒
1.
Wang J  Zhou W  Yuan H  Wang Y 《Carbohydrate research》2008,343(15):2583-2588
A 28kDa chitosanase designated as Csn2 was purified from the culture broth of the fungus Gongronella sp. JG through three chromatography steps: CM-Sepharose FF, Superdex 200 and SP-Sepharose FF. Its optimal reaction pH and temperature were pH 5.6 and between 55 degrees C and 60 degrees C. The half-lives of Csn2 at 50 degrees C and 55 degrees C were estimated to be 30min and 11min, respectively. The K(m) value of Csn2 in sodium acetate buffer (pH 5.6) at 55 degrees C was 8.86mg/mL. Mn(2+), Ca(2+) and Sr(2+) were activators of Csn2; ETDA was an inhibitor. Cu(2+) stimulated Csn2 at 1mM, but inhibited Csn2 activity at 10mM. Csn2 displayed strong activity on colloidal chitosan, but did not hydrolyze colloidal chitin and carboxylmethyl cellulose. Thin layer chromatography analysis showed the end products of colloidal chitosan hydrolyzed by Csn2 were chitobiose, chitotriose and chitotetraose with chitotriose as the major product. The N terminus of Csn2 was determined to be YQLPANLKKIYDSHKSGTC. Part of the genomic DNA sequence corresponding to Csn2 was cloned. Sequence alignment showed DNA sequence of Csn2 was partly identical to chitosanase genes from Metarhizium anisopliae var. acridum, Hypocrea lixii and Aspergillus fumigatus. Based on sequence similarity, Csn2 was classified as a GH-75 chitosanase.  相似文献   

2.
We have taken advantage of the intrinsic fluorescence properties of chitosanases to rapidly and quantitatively evaluate the protective effect of chitosan against thermal denaturation of chitosanases. The studies were done using wild type chitosanases N174 produced by Streptomyces sp. N174 and SCO produced by Streptomyces coelicolor A3(2). In addition, two mutants of N174 genetically engineered by single amino acid substitutions (A104L and K164R) and one "consensus" (N174-CONS) chitosanase designed by multiple amino acid substitutions of N174 were analyzed. Chitosan used had a weight average molecular weight (Mw) of 220 kDa and was 85% deacetylated. Results showed a pH and concentration-dependent protective effect of chitosan in all the cases. However, the extent of thermal protection varied depending on chitosanases, suggesting that key amino acid residues contributed to resistance to heat denaturation. The transition temperatures (T(m)) of N174 were 54 degrees C and 69.5 degrees C in the absence and presence (6 g/l) of chitosan, respectively. T(m) were increased by 11.6 degrees C (N174-CONS), 13.8 degrees C (CSN-A104L), 15.6 degrees C (N174-K164R) and 25.2 degrees C (SCO) in the presence of chitosan (6 g/l). The thermal protective effect was attributed to an enzyme-ligand thermostabilization mechanism since it was not mimicked by the presence of anionic (carboxymethyl cellulose, heparin) or cationic (polyethylene imine) polymers, polyhydroxylated (glycerol, sorbitol) compounds or inorganic salts. Furthermore, the data from fluorometry experiments were in agreement with those obtained by analysis of reaction time-courses performed at 61 degrees C in which case CSN-A104L was rapidly inactivated whereas N174, N174-CONS and N174-K164R remained active over a reaction time of 90 min. This study presents evidence that (1) the fluorometric determination of T(m) in the presence of chitosan is a reliable technique for a rapid assessment of the thermal behavior of chitosanases, (2) it is applicable to structure-function studies of mutant chitosanases and, (3) it can be useful to provide an insight into the mechanism by which mutations can influence chitosanase stability.  相似文献   

3.
Bacillus megaterium P1, a bacterial strain capable of hydrolyzing chitosan, was isolated from soil samples. Chitosan-degrading activity was induced by chitosan but not by its constituent d-glucosamine. Extracellular secretion of chitosanase reached levels corresponding to 1 U/ml under optimal conditions. Three chitosan-degrading proteins (chitosanases A, B, and C) were purified to homogeneity. Chitosanase A (43 kilodaltons) was highly specific for chitosan and represented the major chitosan-hydrolyzing species. Chitosanases B (39.5 kilodaltons) and C (22 kilodaltons) corresponded to minor activities and possessed comparable specific activities toward chitosan, chitin, and cellulose. Chitosanase A was active from pH 4.5 to 6.5 and was stable on the basis of activity up to 45 degrees C. The optimum temperature for enzymatic chitosan hydrolysis was 50 degrees C. Kinetic studies on chitosanase A suggest that the enzyme is substrate inhibited. The apparent K(m) and V(max) determined at 22 degrees C and pH 5.6 were 0.8 mg/ml and 280 U/mg, respectively. End products of chitosan hydrolysis by each of the three chitosanases were identified as glucosamine oligomers, similar to those obtained for previously reported chitosanase digestions.  相似文献   

4.
Streptomyces griseus HUT 6037 inducibly produced two chitosanases when grown on chitosan. To elucidate the mechanism of degradation of chitinous compound by this strain, chitosanases I and II of S. griseus HUT 6037 were purified and characterized. The purified enzymes had a molecular mass of 34 kDa. Their optimum pH was 5.7, and their optimum temperature was 60 degrees C. They hydrolyzed not only partially deacetylated chitosan, but also carboxymethylcellulose. Time-dependent 1H-NMR spectra showing hydrolysis of (GlcN)6 by the chitosanases were obtained for identification of the anomeric form of the reaction products. Both chitosanases produced the beta-form specifically, indicating that they were retaining enzymes. These enzymes catalyzed a glycosyltransfer reaction in the hydrolysis of chitooligosaccharides. The N-terminal and internal amino acid sequences of chitosanase II were identified. A PCR fragment corresponding to these amino acid sequences was used to screen a genomic library for the entire gene encoding chitosanase II. Sequencing of the choII gene showed an open reading frame encoding a protein with 359 amino acid residues. The deduced primary structure was similar to endoglucanase E-5 of Thermomonospora fusca, which enzyme belongs to family 5 of the glycosyl hydrolases. This is the first report of a family 5 chitosanase with transglycosylation activity.  相似文献   

5.
刘怀伟  鲍晓明 《微生物学报》2009,49(12):1607-1612
摘要:【目的】本研究旨在了解腐皮镰孢菌(Fusarium solani)壳聚糖酶的基本酶学性质及其在壳寡糖生产中的应用,构建能高效分泌表达壳聚糖酶的酿酒酵母工业菌株。【方法】采用RT-PCR扩增腐皮镰孢菌壳聚糖酶的cDNA序列;通过组氨酸标签,纯化得到E. coli表达的重组壳聚糖酶,并进行基本酶学性质研究;以薄层层析、高效液相色谱等技术对该酶的酶解产物进行分析;通过马克斯克鲁维酵母(Kluyveromyces marxianus)菊粉酶信号肽(INU1A)实现壳聚糖酶在酿酒酵母工业菌株N-27中的分泌表  相似文献   

6.
The 3D structure-oriented alignment of the primary sequences of fourteen chitosanases, mainly of bacterial origin and belonging to families 46 and 80 of glycoside hydrolases, resulted in the identification of the following pattern common to all these enzymes: E-[DNQ]-x(8,17)-Y-x(7)-D-x-[RD]-[GP]-x-[TS]-x(3)-[AIVFLY]-G- x(5,11)-D. This pattern is proposed as the molecular signature of the chitosanases from families 46 and 80. It includes several amino acids essential for enzyme activity and (or) stability as shown by site-directed mutagenesis studies on the chitosanase from Streptomyces sp. N174. In particular, it includes two carboxylic residues directly involved in catalysis. We suggest that there is a continuum of sequence similarity between all the analyzed chitosanases, and that all these enzymes should probably be classified in one family.  相似文献   

7.

The genome of Kitasatospora setae KM-6054, a soil actinomycete, has three genes encoding chitosanases belonging to GH46 family. The genes (csn1-3) were cloned in Streptomyces lividans and the corresponding enzymes were purified from the recombinant cultures. The csn2 clone yielded two proteins (Csn2BH and Csn2H) differing by the presence of a carbohydrate-binding domain. Sequence analysis showed that Csn1 and Csn2H were canonical GH46 chitosanases, while Csn3 resembled chitosanases from bacilli. The activity of the four chitosanases was tested in a variety of conditions and on diverse chitosan forms, including highly N-deacetylated chitosan or chitosan complexed with humic or polyphosphoric acid. Kinetic parameters were also determined. These tests unveiled the biochemical diversity among these chitosanases and the peculiarity of Csn3 compared with the other three enzymes. The observed biochemical diversity is discussed based on structural 3D models and sequence alignment. This is a first study of all the GH46 chitosanases produced by a single microbial strain.

  相似文献   

8.
For the enzymatic production of chitosan oligosaccharides from chitosan, a chitosanase-producing bacterium, Bacillus sp. strain KCTC 0377BP, was isolated from soil. The bacterium constitutively produced chitosanase in a culture medium without chitosan as an inducer. The production of chitosanase was increased from 1.2 U/ml in a minimal chitosan medium to 100 U/ml by optimizing the culture conditions. The chitosanase was purified from a culture supernatant by using CM-Toyopearl column chromatography and a Superose 12HR column for fast-performance liquid chromatography and was characterized according to its enzyme properties. The molecular mass of the enzyme was estimated to be 45 kDa by means of sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme demonstrated bifunctional chitosanase-glucanase activities, although it showed very low glucanase activity, with less than 3% of the chitosanase activity. Activity of the enzyme increased with an increase of the degrees of deacetylation (DDA) of the chitosan substrate. However, the enzyme still retained 72% of its relative activity toward the 39% DDA of chitosan, compared with the activity of the 94% DDA of chitosan. The enzyme produced chitosan oligosaccharides from chitosan, ranging mainly from chitotriose to chitooctaose. By controlling the reaction time and by monitoring the reaction products with gel filtration high-performance liquid chromatography, chitosan oligosaccharides with a desired oligosaccharide content and composition were obtained. In addition, the enzyme was efficiently used for the production of low-molecular-weight chitosan and highly acetylated chitosan oligosaccharides. A gene (csn45) encoding chitosanase was cloned, sequenced, and compared with other functionally related genes. The deduced amino acid sequence of csn45 was dissimilar to those of the classical chitosanase belonging to glycoside hydrolase family 46 but was similar to glucanases classified with glycoside hydrolase family 8.  相似文献   

9.
Pseudomonas sp. A-01, isolated as a strain with chitosan-degrading activity, produced a 28 kDa chitosanase. Following purification of the chitosanase (Cto1) and determination of its N-terminal amino acid sequence, the corresponding gene (cto1) was cloned by a reverse-genetic technique. The gene encoded a protein, composed of 266 amino acids, including a putative signal sequence (1-28), that showed an amino acid sequence similar to known family-46 chitosanases. Cto1 was successfully overproduced and was secreted by a Brevibacillus choshinensis transformant carrying the cto1 gene on expression plasmid vector pNCMO2. The purified recombinant Cto1 protein was stable at pH 5-8 and showed the best chitosan-hydrolyzing activity at pH 5. Replacement of two acidic amino acid residues, Glu23 and Asp41, which correspond to previously identified active centers in Streptomyces sp. N174 chitosanase, with Gln and Asn respectively caused a defect in the hydrolyzing activity of the enzyme.  相似文献   

10.
beta-mannanase (EC 3.2.1.78) from Bacillus subtilis SA-22 was purified successively by ammonium sulfate precipitation, hydroxyapatite chromatography, Sephadex G-75 gel filtration and DEAE-52 anion-exchange chromatography. Through these steps, the enzyme was concentrated 30.75-fold with a recovery rate of 23.43%, with a specific activity of 34780.56 u/mg. Molecular weight of the enzyme was determined to be 38 kD by SDS-PAGE and 34 kD by gel filtration. The results revealed that the optimal pH value for the enzyme was 6.5 and the optimal temperature was 70 degrees C. The enzyme is stable between pH 5 to 10. The enzyme remained most of its activity after a treatment of 4 h at 50 degrees C, but lost 25% of activity at 60 degrees C for 4 h, lost 50% of activity at 70 degrees C for 3 h. The enzyme activity was strongly inhibited by Hg2+. The Michaelis constants (Km) were measured as 11.30 mg/mL for locust bean gum and 4.76 mg/mL for konjac powder, while Vmax for these two polysaccharides were 188.68 (micromol x mL(-1) x min(-1)) and 114.94 (micromol x mL(-1) x min(-1)), respectively.  相似文献   

11.
4-Methylumbelliferyl beta-chitotrioside [(GlcN)(3)-UMB] was prepared from 4-methylumbelliferyl tri-N-acetyl-beta-chitotrioside [(GlcNAc)(3)-UMB] using chitin deacetylase from Colletotrichum lindemuthianum, and hydrolyzed by chitosanase from Streptomyces sp. N174. The enzymatic deacetylation of (GlcNAc)(3)-UMB was confirmed by (1)H-NMR spectroscopy and mass spectrometry. When the (GlcN)(3)-UMB obtained was incubated with chitosanase, the fluorescence intensity at 450 nm obtained by excitation at 360 nm was found to increase with proportion to the reaction time. The rate of increase in the fluorescence intensity was proportional to the enzyme concentration. This indicates that chitosanase hydrolyzes the glycosidic linkage between a GlcN residue and UMB moiety releasing the fluorescent UMB molecule. Since (GlcN)(3) itself cannot be hydrolyzed by the chitosanase, (GlcN)(3)-UMB is considered to be a useful low molecular weight substrate for the assay of chitosanase. The k(cat) and K(m) values obtained for the substrate (GlcN)(3)-UMB were determined to be 8.1 x 10(-5) s(-1) and 201 microM, respectively. From TLC analysis of the reaction products, the chitosanase was found to hydrolyze not only the linkages between a GlcN residue and UMB moiety, but also the linkages between GlcN residues. Nevertheless, the high sensitivity of the fluorescence detection of the UMB molecule would enable a more accurate determination of kinetic constants for chitosanases.  相似文献   

12.
We have investigated the mechanism of the interaction of Streptomyces sp. N174 chitosanase with glucosamine hexasaccharide [(GlcN)(6)] by site-directed mutagenesis, thermal unfolding, and (GlcN)(6) digestion experiments, followed by theoretical calculations. From the energy-minimized model of the chitosanase-(GlcN)(6) complex structure (Marcotte et al., 1996), Asp57, which is present in all known chitosanases, was proposed to be one of the amino acid residues that interacts with the oligosaccharide substrate. The chitosanase gene was mutated at Asp57 to Asn (D57N) and Ala (D57A), and the relative activities of the mutated chitosanases were found to be 72 and 0.5% of that of the wild type, respectively. The increase in the transition temperature of thermal unfolding (T(m)), usually observed upon the addition of (GlcN)(n) to chitosanase mutants unaffected in terms of substrate binding, was considerably suppressed in the D57A mutant. These data suggest that Asp57 is important for substrate binding. The experimental time-courses of [(GlcN)(6)] degradation were analyzed by a theoretical model in order to obtain the binding free energy values of the individual subsites of the chitosanases. A (-3, -2, -1, +1, +2, +3) subsite model agreed best with the experimental data. This analysis also indicated that the mutation of Asp57 affects substrate affinity at subsite (-2), suggesting that Asp57 most likely participates in the substrate binding at this subsite.  相似文献   

13.
Characterization and kinetics of 45 kDa chitosanase from Bacillus sp. P16   总被引:1,自引:0,他引:1  
An extracellular 45 kDa endochitosanase was purified and characterized from the culture supernatant of Bacillus sp. P16. The purified enzyme showed an optimum pH of 5.5 and optimum temperature of 60 degrees C, and was stable between pH 4.5-10.0 and under 50 degrees C. The Km and Vmax were measured with a chitosan of a D.A. of 20.2% as 0.52 mg/ml and 7.71 x 10(-6) mol/sec/mg protein, respectively. The enzyme did not degrade chitin, cellulose, or starch. The chitosanase digested partially N-acetylated chitosans, with maximum activity for 15-30% and lesser activity for 0-15% acetylated chitosan. The chitosanase rapidly reduced the viscosity of chitosan solutions at a very early stage of reaction, suggesting the endotype of cleavage in polymeric chitosan chains. The chitosanase hydrolyzed (GlcN)7 in an endo-splitting manner producing a mixture of (GlcN)(2-5). Time course studies showed a decrease in the rate of substrate degradation from (GlcN)7 to (GlcN)6 to (GlcN)5, as indicated by the apparent first order rate constants, k1 values, of 4.98 x 10(-4), 2.3 x 10(-4), and 9.3 x 10(-6) sec(-1), respectively. The enzyme hardly catalyzed degradation of chitooligomers smaller than the pentamer.  相似文献   

14.
We have previously shown that sex and obesity independently affect basal very low density lipoprotein (VLDL)-triglyceride (TG) kinetics. In the present study, we investigated the effect of hyperglycemia-hyperinsulinemia on VLDL-TG kinetics in lean and obese men and women (n = 6 in each group). VLDL-TG kinetics were measured during basal, postabsorptive conditions and during glucose infusion (5.5 mg x kg FFM(-1) x min(-1)) by using [(2)H(5)]glycerol bolus injection in conjunction with compartmental modeling analysis. Basal VLDL-TG secretion in plasma was greater in obese than in lean men (7.8 +/- 0.6 and 2.9 +/- 0.4 micromol x l plasma(-1) x min(-1); P < 0.001) but was not different in lean and obese women (5.0 +/- 1.1 and 5.9 +/- 1.1 micromol x l plasma(-1) x min(-1)). Glucose infusion decreased the VLDL-TG secretion rate by approximately 50% in lean and obese men and in lean women (to 1.5 +/- 0.4, 4.0 +/- 0.6, and 2.2 +/- 0.4 micromol x l plasma(-1) x min(-1), respectively; all P < 0.05) but had no effect on the VLDL-TG secretion rate in obese women (4.9 +/- 1.0 micromol x l plasma(-1) x min(-1)). These results demonstrate that both sex and adiposity affect the regulation of VLDL-TG metabolism. Glucose and insulin decrease VLDL-TG production in both lean men and lean women; obesity is associated with resistance to the glucose- and insulin-mediated suppression of VLDL-TG secretion in women, but not in men.  相似文献   

15.
We investigated the uptake of biphenyl by the psychrotolerant, polychlorinated biphenyl (PCB)-degrader, Pseudomonas sp. strain Cam-1 and the mesophilic PCB-degrader, Burkholderia sp. strain LB400. The effects of growth substrates, metabolic inhibitors, and temperature on [14C]biphenyl uptake were studied. Biphenyl uptake by both strains was induced by growth on biphenyl, and was inhibited by dinitrophenol (DNP) and carbonyl cyanide m-chlorophenylhydrazone (CCCP), which are metabolic uncouplers. The Vmax and Km for biphenyl uptake by Cam-1 at 22 degrees C were 5.4 +/- 1.7 nmol x min(-1) x (mg of cell protein)(-1) and 83.1 +/- 15.9 micromol x L(-1), respectively. The Vmax and Km for biphenyl uptake by LB400 at 22 degrees C were 3.2 +/- 0.3 nmol x min(-1) x (mg of cell protein(-1)) and 51.5 +/- 9.6 micromol x L(-1), respectively. At 15 degrees C, the maximum rate for biphenyl uptake by Cam-1 and LB400 was 3.1 +/- 0.3 nmol x min(-1) x (mg of cell protein)(-1) and 0.89 +/- 0.1 nmol x min(-1) x (mg of cell protein)(-1), respectively. Thus, the maximum rate for biphenyl uptake by Cam-1 at 15 degrees C was more than 3 times higher than that for LB400.  相似文献   

16.
In the present study, we explored the effect of the progestin medrogestone on the sulfatase and sulfotransferase activities in the hormone-dependent MCF-7 and T-47D human breast cancer cell lines. After 24 h incubation at 37 degrees C of physiological concentrations of estrone sulfate ([3H]-E1S: 5x10(-9) mol/l), it was observed that this estrogen was converted in a great proportion to E2 in both cell lines. Medrogestone significantly inhibits this transformation, at all the concentrations tested (5x10(-8) to 5x10(-5) mol/l), in both cell lines. The IC50 values were 1.93 micromol/l and 0.21 micromol/l in MCF-7 and T-47D cells, respectively. In another series of studies, after 24 h incubation at 37 degrees C of physiological concentrations of estrone ([3H]-E1: 5x10(-9) mol/l), the sulfotransferase activity was detectable in both cell lines. Estrogen sulfates (ES) are found exclusively in the culture medium, which suggests that as soon as they are formed they are excreted into the medium. Medrogestone has a biphasic effect on sulfotransferase activity in both cell lines. At low doses: 5x10(-8) and 5x10(-7) mol/l, this compound stimulates the enzyme by +73.5 and 52.7%, respectively, in MCF-7, and by 84.5 and 62.6% in T-47D cells. At high concentrations: 5x10(-6) and 5x10(-5) mol/l, medrogestone has no effect on MCF-7 cells, but inhibits the sulfotransferase activity in T-47D cells by -31.4% at 5x10(-5) mol/l. In conclusion, the inhibitory effect provoked by medrogestone on the enzyme involved in the biosynthesis of E2 (sulfatase pathway) in estrogen-dependent breast cancer, as well as the stimulatory effect on the formation of the inactive ES, support a probable anti-proliferative effect of this progestin in breast tissue. Clinical applications of these findings can open new therapeutic possibilities for this disease.  相似文献   

17.
Free fatty acids (FFA) have been shown to inhibit insulin suppression of endogenous glucose production (EGP). To determine whether this is the result of stimulation by FFA of gluconeogenesis (GNG) or glycogenolysis (GL) or a combination of both, we have determined rates of GNG and GL (with (2)H(2)O) and EGP in 16 healthy nondiabetic volunteers (11 males, 5 females) during euglycemic-hyperinsulinemic (~450 pM) clamping performed either with or without simultaneous intravenous infusion of lipid plus heparin. During insulin infusion, FFA decreased from 571 to 30 micromol/l (P < 0.001), EGP from 15.7 to 2.0 micromol x kg(-1) x min(-1) (P < 0.01), GNG from 8.2 to 3.7 micromol x kg(-1). min(-1) (P < 0.05), and GL from 7.4 to -1.7 micromol x kg(-1). min(-1) (P < 0.02). During insulin plus lipid/heparin infusion, FFA increased from 499 to 1,247 micromol/l (P < 0.001). EGP decreased 64% less than during insulin alone (-5.1 +/- 0.7 vs. -13.7 +/- 3.4 micromol x kg(-1). min(-1)). The decrease in GNG was not significantly different from the decrease of GNG during insulin alone (-2.6 vs. -4.5 micromol x kg(-1). min(-1), not significant). In contrast, GL decreased 66% less than during insulin alone (-3.1 vs. -9.2 micromol x kg(-1). min(-1), P < 0.05). We conclude that insulin suppressed EGP by inhibiting GL more than GNG and that elevated plasma FFA levels attenuated the suppression of EGP by interfering with insulin suppression of GL.  相似文献   

18.
In this study, the response of the sarcoplasmic reticulum (SR) to prolonged exercise, performed in normoxia (inspired O(2) fraction = 0.21) and hypoxia (inspired O(2) fraction = 0.14) was studied in homogenates prepared from the vastus lateralis muscle in 10 untrained men (peak O(2) consumption = 3.09 +/- 0.25 l/min). In normoxia, performed at 48 +/- 2.2% peak O(2) consumption, maximal Ca(2+)-dependent ATPase activity was reduced by approximately 25% at 30 min of exercise compared with rest (168 +/- 10 vs. 126 +/- 8 micromol.g protein(-1) x min(-1)), with no further reductions observed at 90 min (129 +/- 6 micromol x g protein(-1) x min(-1)). No changes were observed in the Hill coefficient or in the Ca(2+) concentration at half-maximal activity. The reduction in maximal Ca(2+)-dependent ATPase activity at 30 min of exercise was accompanied by oxalate-dependent reductions (P < 0.05) in Ca(2+) uptake by approximately 20% (370 +/- 22 vs. 298 +/- 25 micromol x g protein(-1) x min(-1)). Ca(2+) release, induced by 4-chloro-m-cresol and assessed into fast and slow phases, was decreased (P < 0.05) by approximately 16 and approximately 32%, respectively, by 90 min of exercise. No differences were found between normoxia and hypoxia for any of the SR properties examined. It is concluded that the disturbances induced in SR Ca(2+) cycling with prolonged moderate-intensity exercise in human muscle during normoxia are not modified when the exercise is performed in hypoxia.  相似文献   

19.
Pseudomonas sp. A-01, isolated as a strain with chitosan-degrading activity, produced a 28 kDa chitosanase. Following purification of the chitosanase (Cto1) and determination of its N-terminal amino acid sequence, the corresponding gene (cto1) was cloned by a reverse-genetic technique. The gene encoded a protein, composed of 266 amino acids, including a putative signal sequence (1-28), that showed an amino acid sequence similar to known family-46 chitosanases. Cto1 was successfully overproduced and was secreted by a Brevibacillus choshinensis transformant carrying the cto1 gene on expression plasmid vector pNCMO2. The purified recombinant Cto1 protein was stable at pH 5–8 and showed the best chitosan-hydrolyzing activity at pH 5. Replacement of two acidic amino acid residues, Glu23 and Asp41, which correspond to previously identified active centers in Streptomyces sp. N174 chitosanase, with Gln and Asn respectively caused a defect in the hydrolyzing activity of the enzyme.  相似文献   

20.
There is a growing interest in chitosanases as enzymatic tools to hydrolyze chitosan into bioactive forms: low molecular weight chitosan (LMWC) or chitosan oligosaccharides (CHOS). However chitosanases are still expensive and methods of large-scale production of these enzymes are not yet established. The article reviews the approaches used for chitosanase production in various bacterial hosts, pointing out the difficulties resulting from the necessity to include chitosan into the medium composition. A mutated Streptomyces host allows for the efficient production of several chitosanases originating from actinobacteria in the absence of chitosan as inducer.  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