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.
Chitinases I and II were purified from the culture supernatant of Aeromonas sp. 10S-24 by ammonium sulfate precipitation, SP-Sephadex C-50 chromatography, Sephacryl S-200 gel filtration, and chromatofocusing. Both enzymes were most active at pH 4.0 and the optimum temperature for I and II were 50°C and 60°C. Chitinase I was stable at pHs between 4 and 9 and at temperatures below 50°C and chitinase II was stable at pHs between 5 and 7 and at temperatures below 45°C. The molecular weights were estimated by 8D8 polyacrylamide gel electrophoresis to be 112,000 and 115,000 for I and II respectively, while gel filtration showed the molecular weight to be 114,000 for both types of the enzyme. The pIs for I and II were 7.9 and 8.1, respectively. The activities of both enzymes were inhibited by Ag+ and iodoacetic acid.  相似文献   

2.
Two lytic enzymes (enzyme I and enzyme II) that lysed Micrococcus lysodeikticus were isolated from the crude extract of Polysphondylium pallidum myxamoebae grown in the presence of Klebsiella aerogenes by precipitation with protamine sulfate and by chromatography on DEAE-Sepharose CL-6B. Enzyme I was further purified by gel filtration on a Superose12 column, and enzyme II by chromatography on a MonoQ HR 5/5 column and gel filtration on a Superose12 column. Enzyme I was a basic protein, while enzyme II was acidic. The molecular weights of enzyme I and II were about 14,000 and 22,000, respectively by SDS-polyacrylamide gel electrophoresis. Optimum pHs for the activity were 5.0 for enzyme I and between 3.5 and 4.0 for enzyme II. The maximum activity of enzyme I and II was obtained at 65°C and 45°C to 55°C and at ionic strength of 0.0075 to 0.03 and 0.06, respectively. Both enzymes cleaved the glycosidic bond of β(1,4)-N-acetylmuramyl-acetylglucosamine of the cell wall peptidoglycan of Micrococcus lysodeikticus. These results indicate that the two lytic enzymes of Polysphondylium pallidum myxamoebae are N-acetylmuramidases.  相似文献   

3.
Common carp is cheap and prolific in Australian waters and is regarded as an aquatic environmental pest. In order to add value to this fish species, surimi and kamaboko was prepared from common carp and its rheological and microstructure characteristics were compared with those produced from Alaska pollock and threadfin bream. Temperature sweep tests were run under 100-Pa stress and 0.1-Hz frequency, obtained from linear viscoelastic ranges of all tested fish gels. Storage modulus (G′) thermographs of all samples consisted of two positive peaks and a plateau zone in between. The sol-gel transition was completed at about 53 and 61 °C for these Alaska pollock and threadfin bream gels, respectively, whereas it was recorded at about 69 °C for common carp gel. At these temperatures, G′ of Alaska pollock gel was recorded at 330 kPa, which was 71% and 88% greater than that from threadfin bream and common carp gels, respectively. Furthermore, Alaska pollock and threadfin bream gels had greater gel strength than the gel prepared from common carp surimi. Textural quality could be associated with cross-linking in the gel network. From scanning electron microscopy studies, the microstructure of Alaska pollock gel matrix had about 15,450 polygonal structures per square millimeter with an average area of about 9 μm2. For threadfin bream and common carp gels, the polygonal structures were larger and 12% and 39% fewer, respectively, than those of Alaska pollock gel. However, these results cannot be only attributed to the species variation among tested fish as some other factors such as harvest location, physiological state, handling and processing method, etc. were not considered in this study.  相似文献   

4.
Secondary structure, microstructure, and mechanical properties of heat induced 11S globulin gel were studied to discover their relationships. Heat-induced 11S globulin gel at 80, 90, and 95°C were comprised of 7.2, 16.6, and 23.8% of α-helix; 19.4, 19.5, and 27.5% of β-sheet, and 73.5, 64.5, and 48.6%, of random coil, respectively. This indicated the gel formed at higher temperatures contained more α-helix and less random coil structures. Micrographs of gels heated at 90 and 95°C had a more extended and integral matrix. Gel strength of heat-induced gels at 90 and 95°C were significantly greater than that of 80°C. These data indicated that the increase in α-helix of heat-induced 11S globulin gel have facilitated the establishment of a good gel matrix.  相似文献   

5.
The effects of pH (6.7 or 5.8), protein concentration and the heat treatment conditions (70 or 90 °C) on the physical properties of heat-induced milk protein gels were studied using uniaxial compression, scanning electron microscopy, differential scanning calorimetry, and water-holding capacity measurements. The systems were formed from whey protein isolate (10–15% w/v) with (5% w/v) or without the addition of caseinate. The reduction in pH from 6.7 to 5.8 increased the denaturation temperature of the whey proteins, which directly affected the gel structure and mechanical properties. Due to this increase in the denaturation temperature of the β-lactoglobulin and α-lactalbumin, a heat treatment of 70 °C/30 min did not provide sufficient protein unfolding to form self-supporting gels. However, the presence of 5% (w/v) sodium caseinate decreased the whey protein thermo stability and was essential for the formation of self-supporting gels at pH 6.7 with heat treatment at 70 °C/30 min. The gels formed at pH 6.7 showed a fine-stranded structure, with great rigidity and deformability as compared to those formed at pH 5.8. The latter had a particulate structure and exuded water, which did not occur with the gels formed at pH 6.7. The addition of sodium caseinate led to less porous networks with increased gel deformability and strength but decreased water exudation. The same tendencies were observed with increasing whey protein concentration.  相似文献   

6.
An amidase (EC 3.5.1.4) in branch 2 of the nitrilase superfamily, from the thermophilic strain Geobacillus pallidus RAPc8, was produced at high expression levels (20 U/mg) in small-scale fermentations of Escherichia coli. The enzyme was purified to 90% homogeneity with specific activity of 1,800 U/mg in just two steps, namely, heat-treatment and gel permeation chromatography. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and electron microscopic (EM) analysis of the homogenous enzyme showed the native enzyme to be a homohexamer of 38 kDa subunits. Analysis of the biochemical properties of the amidase showed that the optimal temperature and pH for activity were 50 and 7.0°C, respectively. The amidase exhibited high thermal stability at 50 and 60°C, with half-lives greater than 5 h at both temperatures. At 70 and 80°C, the half-life values were 43 and 10 min, respectively. The amidase catalyzed the hydrolysis of low molecular weight aliphatic amides, with d-selectivity towards lactamide. Inhibition studies showed activation/inhibition data consistent with the presence of a catalytically active thiol group. Acyl transfer reactions were demonstrated with acetamide, propionamide, isobutyramide, and acrylamide as substrates and hydroxylamine as the acyl acceptor; the highest reaction rate being with isobutyramide. Immobilization by entrapment in polyacrylamide gels, covalent binding on Eupergit C beads at 4°C and on Amberlite-XAD57 resulted in low protein binding and low activity, but immobilization on Eupergit C beads at 25°C with cross-linking resulted in high protein binding yield and high immobilized specific activity (80% of non-immobilized activity). Characterization of Eupergit C-immobilized preparations showed that the optimum reaction temperature was unchanged, the pH range was somewhat broadened, and stability was enhanced giving half-lives of 52 min at 70°C and 30 min at 80°C. The amidase has potential for application under high temperature conditions as a biocatalyst for d-selective amide hydrolysis producing enantiomerically pure carboxylic acids and for production of novel amides by acyl transfer.  相似文献   

7.
Two types of extracellular lipases (I and II) from Trichosporon fermentans WU-C12 were purified by acetone precipitation and successive chromatographies on Butyl-Toyopearl 650 M, Toyopearl HW-55F and Q-Sepharose FF. The molecular weight of lipase I was 53 kDa by SDS-polyacrylamide gel electrophoresis (SDS-PAGE) and 160 kDa by gel filtration, while that of lipase II was 55 kDa by SDS-PAGE and 60 kDa by gel filtration. For the hydrolysis of olive oil, the optimum pH and temperature of both the lipases were 5.5 and 35°C, respectively. The lipases showed stable activities after incubation at 30°C for 24 h in a pH range from 4.0 to 8.0. The thermostability of lipase I for 30 min at a reaction pH of 5.5 was up to 40°C, while that of lipase II under the same conditions was up to 50°C. Both lipases could hydrolyze the 1-, 2-, and 3-positions of triolein, and cleave all three ester bonds, regardless of the position in the triglyceride.  相似文献   

8.
Alginate is a biopolymer used in drug formulations and for surgical purposes. In the presence of divalent cations, it forms solid gels, and such gels are of interest for immobilization of cells and enzymes. In this work, we entrapped trypsin in an alginate gel together with a known substrate, N α-benzoyl-l-arginine-4-nitroanilide hydrochloride (l-BAPNA), and in the presence or absence of d-BAPNA, which is known to be a competitive inhibitor. Interactions between alginate and the substrate as well as the enzyme were characterized with transmission electron microscopy, rheology, and nuclear magnetic resonance spectroscopy. The biocatalysis was monitored by spectrophotometry at temperatures ranging from 10 to 42 °C. It was found that at 37 and 42 °C a strong acceleration of the reaction was obtained, whereas at 10 °C and at room temperature, the presence of d-BAPNA leads to a retardation of the reaction rate. The same effect was found when the reaction was performed in a non-cross-linked alginate solution. In alginate-free buffer solution, as well as in a solution of carboxymethylcellulose, a biopolymer that resembles alginate, the normal behavior was obtained; however, with d-BAPNA acting as an inhibitor at all temperatures. A more detailed investigation of the reaction kinetics showed that at higher temperature and in the presence of alginate, the curve of initial reaction rate versus l-BAPNA concentration had a sigmoidal shape, indicating an allosteric behavior. We believe that the anomalous behavior of trypsin in the presence of alginate is due to conformational changes caused by interactions between the positively charged trypsin and the strongly negatively charged alginate.  相似文献   

9.
At 22°C the bioluminescence decay kinetics in the in vitro reaction catalysed by Vibrio harveyi luciferase in the presence of different aldehydes–-nonanal, decanal, tridecanal and tetradecanal did not follow the simple exponential pattern and could be fitted to a two-exponential process. One more principal distinction from the first-order kinetics is the dependence of the parameters on aldehyde concentration. The complex bioluminescence decay kinetics are interpreted in terms of a scheme, where bacterial luciferase is able to perform multiple turnovers using different flavin species to produce light. The initial phase of the bioluminescent reaction appears to proceed mainly with fully reduced flavin as the substrate while the final one results from the involvement of flavin semiquinone in the catalytic cycle.  相似文献   

10.
Biomass (CHN), respiration rate and food uptake were estimated for the larval development ofElminius modestus at three temperatures (12, 18, 24°C). Mean values of dry weight, elemental composition and energy equivalents increased exponentially with the development from nauplius II to VI. Dry weight, elemental composition and energy content exhibited the highest values at 18°C. Respiration rates increased with the larval stages expressed by a power function, but increased logarithmically with the dry weight of the larvae. The cypris larvae showed a reduced respiration rate compared with nauplius VI. The ingestion rate was measured at a concentration of 100 cells ofSkeletonema costatum μl−1. At 12 and 18°C ingestion rates increased exponentially and at 24°C by a logarithmic function. The fittings were used to estimate the energy budget ofE. modestus during larval development. The energy content of the larvae increased during the development from nauplius II to VI by a factor of 21 at 12°C, 25 at 24°C and 31 at 18°C. The estimated energy content of the freshly metamorphosed barnacle is 100 mJ (12°C), 130 mJ (24°C) and 150 mJ (18°C). The assimilation- (A/I) and gross growth efficiencies (K1) increased strongly during the development from nauplius II to VI (A/I: 6–14% in nauplius II to 50–90% in nauplius VI; K1: 4% in nauplius II to 75% in nauplius VI). The net growth efficiency (K2) showed a relatively constant level ranging between 57 and 83%.  相似文献   

11.
Fluorescein mercuric acetate causes the unwinding of DNA and binds to the separated bases. This unwinding process can be followed by measuring absorption changes of this reagent. For untreated calf thymus DNA, the initial rate was very slow, and the shape of the kinetic curve was sigmoidal. When double-strand breaks of DNA were produced by DNase II treatment or sonication, the initial rate increased and the sigmoidal character disappeared. The initial rate was shown to be proportional to the concentration of helix ends. From this relation the rate of unwinding was estimated to be 2.0 base pairs/sec at 1.0 × 10?5M fluorescein mercuric acetate and 25°C. DNase I treatment, which produces single-strand breaks and a smaller number of double-strand breaks, also increased the initial rate. However, this increase was due only to the double-strand breaks, that is, single-strand breaks had no significant effect on the initial rate. Also, uv irradiation increased the initial rate linearly with uv dose, at least up to 2 × 105 erg/mm2, suggesting that this increase is due to photoproducts other than usual pyrimidine dimers. We discuss the usefulness of this kinetic method in structural studies of DNA.  相似文献   

12.
Abstract

The new affinity gel reported in this study was prepared using EUPERGIT C250L as a chromatographic bed material, to which etylenediamine spacer arms were attached to prevent steric hindrance between the matrix and ligand, and to facilitate effective binding of the CA-specific ligand, of the aromatic sulfonamide type for the purification of α-carbonic anhydrases (Cas; EC 4.2.1.1). Indeed, the aminoethyl moieties of the affinity gel were derivatized by reaction with 4-isothiocyanatobenzenesulfonamide, with the formation of a thiourea-based gel, having inhibitory effects against CAs. Both bovine erythrocyte carbonic anhydrase BCA and human (h) erythrocyte CA isoforms I, II (hCA I and II) have been purified from hemolysates, by using this affinity gel. The greatest purification fold and column yields for BCA and for cytosolic (hCA I?+?II) enzymes were of 181-fold (21.07%) and 184-fold (9.49%), respectively. Maximum binding was achieved at 15?°C and I?=?0.3 ionic strength for α-carbonic anhydrases.  相似文献   

13.
Aspergillus kawachii α-amylase [EC 3.2.1.1] I and II were purified from shochu koji extract by DEAE Bio-Gel A ion exchange chromatography, Sephacryl S-300 gel chromatography (pH 3.6), coamino dodecyl agarose column chromatography and Sephacryl S-200 gel chromatography. By gel chromatography on a Sephacryl S-300 column, the molecular weights of the purified α-amylase I and II were estimated to be 104,000 and 66,000, respectively. The isoelectric points of α-amylase I and II were 4.25 and 4.20, respectively. The optimal pH range of α-amylase I was 4.0 to 5.0, and the optimum pH of α-amylase II was 5.0. The optimum temperatures of both α-amylases were around 70°C at pH 5.0. Both α-amylases were stable from pH 2.5 to 6.0 and up to 55°C, retaining more than 90% of the original activities. Heavy metal ions such as Hg2 + and Pb2 + were potent inhibitors for both α-amylases.  相似文献   

14.
Taka-Aki Ono  Norio Murata 《BBA》1979,545(1):69-76
The photosynthetic electron transport and phosphorylation reactions were measured in the room temperature region in the thylakoid membranes prepared from the blue-green alga, Anacystis nidulans. The Arrhenius plot of the Hill reaction with 2,6-dichlorophenolindophenol showed a distinct break of straight lines at 21°C in the membranes from cells grown at 38°C, and at 12°C in those from cells grown at 28°C. The Arrhenius plot of the Hill reaction with ferricyanide showed a break at 13°C in the membranes from cells grown at 38°C, and at 7°C in those from cells grown at 28°C. On the other hand, the Arrhenius plot of the System I reaction with methylviologen as an electron acceptor and 2,6-dichlorophenolindophenol and ascorbate as an electron donor system was composed of a straight line in the membranes from cells grown at 28°C as well as at 38°C. The Arrhenius plot of the System II reaction measured by the ferricyanide reduction mediated by silicotungstate in the presence of 3-(3′,4′-dichlorophenyl)-1,1-dimethylurea also showed a break at 11°C in the membranes from cells grown at 38°C.The Arrhenius plot of the phosphorylation mediated by N-methylphenazonium methylsulfate showed a break at 21°C in the membranes from cells grown at 38°C and at 12°C in those from cells grown at 28°C. The Arrhenius plot of the phosphorylation mediated by the System I reaction showed a break at 24°C in the membranes from cells grown at 38°C.The characteristic features in the Arrhenius plots of the photosynthetic electron transport and phosphorylation reactions are discussed in terms of the transition of physical phase of the thylakoid membrane lipids.  相似文献   

15.
The kinetics of the metarhodopsin (meta) I → metarhodopsin II reaction have been studied by flash photolysis in two different types of preparations of bovine rhodopsin: (i) digitonin-solubilized rod outer segment (ROS) membranes with a molar ratio of phospholipid to rhodopsin of approximately 90, and (ii) digitonin-solubilized phospholipid-free rhodopsin with a molar ratio of phospholipid to rhodopsin of less than 0.2. At 20 °C the kinetics in both preparations are multiexponential, but four terms are required to fit the data with the solubilized membranes, whereas only two are required with the phospholipid-free preparation. Thus, phospholipid removal simplifies the kinetics of the meta I → meta II reaction, but the resulting preparation still does not show first-order kinetics. The ratio of the time constants of these two components with detergent-solubilized phospholipid-free rhodopsin was nearly equal to the values found with ROS particles, rhodopsin-phospholipid recombinants and intact rabbit eyes. This suggests a common origin for these two components in all these preparations and appears to exclude heterogeneity in bound phospholipid as the basis of these two-component kinetics.  相似文献   

16.
The glucocortiocoid receptors in the cytosol of neural retina of the 15-day chick embryo were analyzed by quantitative polyacrylamide gel electrophoresis. Maintenance of the triamcinolone acetonide (TA)-receptor complexes under conditions of electrophoretic analysis is dependent on temperatures not exceeding ?2 °C and is favored by low ionic strength, but is relatively insensitive to changes in pH between 5 and 10. Polyacrylamide gel electrophoresis in highly crosslinked Resolving Gels (15% crosslinking with N,N′-diallyltartardiamide) at low wattage and under temperature control at ?2 °C, allowed for detection and partial characterization of over 80% of the specific TA-binding activity of the tissue. One form of the glucocorticoid receptor, designated as complex II, was found to have a molecular weight (Mr) of 175,000. In addition, specifically bound TA was found in a multimillion Mr aggregate which was unable to enter gels of any concentration investigated and has been designated TA-complex I. The ratio of complex I/complex II increased with increasing gel concentration, indicating physical or chemical interaction between II and I. A polyacrylamide gel electrophoresis rerun of isolated TA-complex II gave rise to two smaller TA-binding species: Component B, of Mr 108,000 and component A, a relatively fast migrating molecule which could not be characterized under the conditions used. The ratio of BA appeared constant and close to 2, suggesting that A and B may be significant structural elements of complex II. Polyacrylamide gel electrophoresis of isolated TA-complex I gave rise to component C of Mr 60,000, but not to components A or B. Components A and B associated to a large Mr complex, designated as I′, which was revealed to an extent directly proportional to gel concentration. Similarly, component C aggregated to I″, as evidenced at elevated gel concentrations. In conclusion, it has been possible to define by gel electrophoresis three of the molecular species (A, B, and C) that comprise the glucocorticoid receptor, and the possible relationships between them.  相似文献   

17.
Creatine kinase thermal aggregation kinetics has been studied in 30 mM Hepes-NaOH buffer, pH 8.0, at two temperatures: 50.6 and 60°C. Aggregation kinetics was analyzed by measuring the growth of apparent absorption (A) at 400 nm. It was found that the limiting value of apparent absorption (A lim) is proportional to protein concentration at both temperatures. The first order rate constant (k I) does not depend on protein concentration in the range 0.05–0.2 mg/ml at temperature 50.6°C, but at temperature 60°C it increases with the growth of protein concentration in the range 0.1–0.4 mg/ml. Kinetic curves, shown in coordinates {A/A lim; t}, in experiments at 50.6°C fuse to a common curve, which coincides with the theoretical curve of creatine kinase denaturation calculated using the denaturation rate constant determined from differential scanning calorimetry. At temperature 60°C, half-transformation time t 1/2 = ln2/k I decreases when protein concentration grows. We conclude that when temperature increased from 50.6 to 60°C, change in the kinetic regime of thermal creatine kinase aggregation took place: at 50.6°C aggregation rate is limited by the stage of protein molecule denaturation, but at 60°C it is limited by the stage of protein aggregate growth, which proceeds as a reaction of pseudo-first order. Small heat shock protein Hsp 16.3 Mycobacterium tuberculosis suppresses the creatine kinase aggregation. Published in Russian in Biokhimiya, 2006, Vol. 71, No. 3, pp. 408–416.  相似文献   

18.
The kinetics of the hydrogen-deuterium exchange reactions of double-helical poly (rI) · poly (rC), single-stranded poly(rC) and poly(rI), inosine, and cytosine- 5′-phosphoric acid have been examined, at various temperatures in the range 20 °C to 52 °C, by stopped-flow ultraviolet spectrophotometry, in the region 270 to 300 nm. For the solution of double-helical poly(rI) · poly(rC), two first-order deuteration reactions were found: a fast one and a slow one. At 25 °C and at pH 7.0, the rate constant was 12.3 s?1 for the fast reaction, and 0.13 s?1 for the slow reaction. The rate constant of the fast reaction is nearly equal to that of the single-stranded poly(rC) (12.6 s?1), and is assigned to the deuteration at the amino hydrogen (that is, free from the C · I hydrogen bond) of the cytosine residue. The slow reaction is attributable to the deuteration of the two hydrogens: the amino hydrogen of rC and imide hydrogen of rI, which are rapidly exchanging with each other within every rC · rI base-pair. From the observed temperature effect on this slow reaction rate, it has been concluded that there are two types of “opening process” that are relevant to the hydrogen exchange reaction; one of them is predominent in the range 47 °C to 52 °C and the other in the temperature region lower than 47 °C. The enthalpy (H) and entropy (S) differences of the “open” and “closed” forms in the former type process are ΔH = 167 kcal per mole and ΔS = 507 e.u., while in the latter ΔH = 8.1 kcal per mole and ΔS = 10 e.u..  相似文献   

19.
The metastable state silk I structures of Bombyx mori silk fibroin in the solid state were studied on the basis of 15N- and 13C-nmr chemical shifts of Ala, Ser, and Gly residues. The 15N cross-polarization magic angle spinning (CP/MAS) nmr spectra of the precipitated fraction after chymotrypsin hydrolysis of B. mori silk fibroin with the silk I and silk II forms were measured to determine the 15N chemical shifts of Gly, Ala, and Ser residues. For comparison, 15N CP/MAS nmr chemical shifts of Ala were measured for [15N] Ala Philosamia cynthia ricini silk fibroin with antiparallel β-sheet and α-helix forms. The 13C CP/MAS nmr chemical shifts of Ala, Ser, and Gly residues of B. mori silk fibroin with the silk I and silk II forms, as well as 13C CP/MAS nmr chemical shifts of Ala residue of P. c. ricini silk fibroin with β-sheet and α-helix forms, are used for the examination of the silk I structure. Both silk I and α-helix peaks are shifted to a lower field than silk II (β-sheet) for the Cα carbons of the Ala residues, while both Cβ carbon peaks are shifted to higher field. However, the silk I peak of the 15N nucleus of the Ala residue is shifted to lower field than the silk II peak, but the α-helix peak is shifted to high field. Thus, the difference in the structure between the silk I and α-helix is reflected in a different manner between the 13C and 15N chemical shifts. The Cα and Cβ chemical shift contour plots for Ala and Ser residues, and the Cα plot for the Gly residue, were prepared from the Protein Data Bank data obtained for 12 proteins and used for discussing the silk I structure quantitatively from the conformation-dependent chemical shifts. The plots reported by Le and Oldfield for 15N chemical shifts were also used for the purpose. All these chemical shift data support Fossey's model (Ala: ϕ = −80°, φ = 150°, Gly: ϕ = −150°, φ = 80°) and do not support Lotz and Keith's model (Ala: ϕ = −104.6°, φ = 112.2°, Gly: ϕ = 79.8°, φ = 49.7° or Ala: ϕ = −124.5°, φ = 88.2°, Gly: ϕ = −49.8°, φ = −76.1°) as the silk I structure. © 1997 John Wiley & Sons, Inc.  相似文献   

20.
An isomaltotriose-producing dextranase II, detected in the culture supernatant of Flavobacterium sp. M-73, was purified to an electrophoretically pure state. Successive chromatography on hydrophobic columns of Amberlite CG-50 and aminooctyl-Sepharose was very effective as the first step of purification. Further purification of the enzyme was performed by affinity column chromatography on isomaltotriose-Sepharose and preparative polyacrylamide gel electrophoresis.

The purified enzyme was shown to be a monomer and had a molecular weight of 114,000. Dextranase II was most active at pH 7.0 and 35°C. It was stable at 4°C for 24 hr over a pH range of 6.5~12.0 and up to 35°C on heating for 10 min. This enzyme had a strict specificity for consecutive α-l,6-glucosidic linkages and readily hydrolyzed clinical dextran and Sephadex gels. The degree of hydrolysis of clinical dextran was 31% expressed as apparent conversion into D-glucose. The amount of isomaltotriose in the hydrolyzate was determined to be 63%.  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