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.
Since there is evidence that oxalyl thiolesters (RSCOCOO) are present in animal cells, and possibly may participate in the control of metabolism, the present study was undertaken to characterize their reactivity with nucleophiles so that one could gain a better understanding of how they might be affecting the activities of enzymes. At 25°C and neutral pH, N-acetyl-S-oxalyl-2-aminoethanethiol (NAC-S-Ox) reacts rapidly with cysteamine (2-aminoethanethiol) to give N-acetylcysteamine and N-oxalylcysteamine. Under similar conditions, other aminothiols, such as cysteine, homocysteine, penicillamine, and cysteine ethyl ester, also react rapidly with NAC-S-Ox, but non-thiol-containing amines, such as alanine, alanine ethyl ester, glycine, and S-methylcysteine, react more than four orders of magnitude less rapidly. The aminothiol reactions apparently proceed by rate-determining oxalyl transfer to the thiol followed by a rapid intramolecular S- to N-oxalyl migration. The reactions follow second-order kinetics with the thiolate anion being the reactive nucleophile. At 25°C and ionic strength 1.0 , kN, defined in the equation, rate = kN[RS][NAC-S-Ox], has the following values ( −1 s−1) for the anion of the reacting thiol: cysteamine, 170; cysteine, 260; cysteine ethyl ester, 76; homocysteine, 380. Rate data for the reaction of NAC-S-Ox with hydroxylamine, imidazole, hydroperoxide, and hydroxide were also obtained. The reaction of S-oxalyl-p-thiocresol with thiol anions under the same conditions gives the following values for kN ( −1 s−1 × 10−3): glutathione, 5.6; N-acetylcysteamine, 3.7; pantetheine, 4.8; 8-mercaptooctanoic acid, 4.5; 6-mercaptooctanoic acid, 1.0; dihydrolipoic acid, 8.2. These results indicate that oxalyl transfers from oxalyl thiolesters to thiol anions occur more than two orders of magnitude more rapidly than corresponding acetyl transfers, and that under physiological conditions any in vivo oxalyl thiolester would equilibrate within minutes with virtually every thiol in the cell, including those attached to enzymes. Consequently, it is proposed that one mechanism by which oxalyl thiolesters may function in vivo to alter the catalytic activities of enzymes is to covalently modify enzymic thiols by acylation with an oxalyl group.  相似文献   

2.
A comparative study of different derivatization procedures has been performed in order to improve the stability of the reaction products o-phthalaldehyde–N-acetylcysteine (OPA–NAC) polyamines. Procedures such as solution derivatization, solution derivatization followed by retention on a packing support, derivatization on different packing supports and on-column derivatization, have been optimized and compared. The degradation rate constant (k) of the derivative was dependent on the procedure used and on the analyte. For the spermine (the most unstable isoindol tested) k was 8±2×10−2 min−1 in solution versus 7.7±1.1×10−4 min−1 on the (C18) solid support. The results obtained showed that forming the derivative on the packing support (C18) gave the best results following this procedure: conditioning the cartridges with borate buffer (1 ml, 0.5 M, pH 8), retention of the analyte, addition of 0.8 ml of OPA–NAC reagent, 0.2 ml borate buffer 0.8 M (pH 8) and elution of the isoindol with 3 ml of MeOH–borate buffer (9:1). The different derivatization procedures have been used to study the stability of the reaction products OPA–NAC polyamines formed in urine matrix using spermine as model compound. Similar results were obtained for standard solutions and urine samples.  相似文献   

3.
Human α1-antitrypsin (AAT) was produced in the recombinant yeast Saccharomyces cerevisiae ATCC 20699 grown in batch and fed-batch culture. The final biomass concentration and antitrypsin concentration attained were 55 g·L−1 and 1.23 g·L−1, respectively, in the fed-batch. The maximum productivities of biomass and antitrypsin were 1.6 and > 0.04 g L−1h−1, respectively, or substantially greater than the highest productivity values reported in the past. For recovering the antitrypsin, the cell slurry was concentrated 4-fold (231 g·L−1 biomass, 122 min of processing) by cross-flow microfiltration and the cells were disrupted by bead milling (3 passes of 3 min total retention time). The cell homogenate was treated with aluminum chloride or PBS (pH 7) to aid separation of the cell debris by flocculation and sedimentation. The clarified cell homogenate was subjected to ammonium sulfate fractionation to precipitate the recombinant antitrypsin. The AAT precipitated at 45–75% saturation of ammonium sulfate, depending on the age of the homogenate. The crude AAT in the homogenate degraded at room temperature (25°C), with a zero order deactivation rate of 1.815 × 10−3 ± 3.43 × 10−4 g AAT L−1h−1.  相似文献   

4.
A hydrogen peroxide permselective membrane with asymmetric structure was prepared and -glucose oxidase (EC 1.1.3.4) was immobilized onto the porous layer. The activity of the immobilized -glucose oxidase membrane was 0.34 units cm−2 and the activity yield was 6.8% of that of the native enzyme. Optimum pH, optimum temperature, pH stability and temperature stability were found to be pH 5.0, 30–40°C, pH 4.0–7.0 and below 55°C, respectively. The apparent Michaelis constant of the immobilized -glucose oxidase membrane was 1.6 × 10−3 mol l−1 and that of free enzyme was 4.8 × 10−2 mol l−1. An enzyme electrode was constructed by combination of a hydrogen peroxide electrode with the immobilized -glucose oxidase membrane. The enzyme electrode responded linearly to -glucose over the concentration 0–1000 mg dl−1 within 10 s. When the enzyme electrode was applied to the determination of -glucose in human serum, within day precision (CV) was 1.29% for -glucose concentration with a mean value of 106.8 mg dl−1. The correlation coefficient between the enzyme electrode method and the conventional colorimetric method using a free enzyme was 0.984. The immobilized -glucose oxidase membrane was sufficiently stable to perform 1000 assays (2 to 4 weeks operation) for the determination of -glucose in human whole blood. The dried membrane retained 77% of its initial activity after storage at 4°C for 16 months.  相似文献   

5.
The kinetics of fungal peroxidase-catalyzed phenolic compounds (PCs) oxidation was investigated in presence of acetylenic-based surfactant Dynol 604 at pH 5.5 and 25 °C. It was shown that the presence of ppm concentrations of surfactant did not influence initial rate of PCs oxidation. The calculated apparent bimolecular rate constants were (1.8 ± 0.2) × 105 M−1 s−1, (1.4 ± 0.4) × 107 M−1 s−1, (1.30 ± 0.06) × 107 M−1 s−1 and 1.1 × 108 M−1 s−1 for phenol, 1-naphthol, 2-naphthol and 1-hydroxypyrene, respectively.During an extensive substrates conversion Dynol 604 showed diverse action for different PCs. The oxidation of phenol practically did not change, whereas the surfactant enhanced the conversion of 1- and 2-naphthol and 1-hydroxypyrene in dose response manner. The results accounted by a scheme, which contains a stadium of enzyme inhibition by oligomeric PC oxidation products. The action of the surfactant was explained by avoidance the enzyme active center clothing with the oligomers. The results acquired demonstrate a remarkable increase of substrates conversion in the presence of Dynol 604.  相似文献   

6.
A polygalacturonase with a molecular mass of 74 kDa, an isoelectric point around pH 4.2 and pH – and temperature optima of 3.9 and 50°C, respectively, was purified from a culture fluid of Penicillium frequentans. The enzyme was characterized as an exo-α-1,4-polygalacturonase (exo-PG I). Km and Vmax for sodium polypectate hydrolysis were 0.68 g/l and 596.8 U × mg−1, respectively. The enzyme, a glycoprotein with a carbohydrate content of 81%, is probably the main pectinase of Penicillium frequentans responsible for cleaving monomer units from the non-reducing end of pectin.  相似文献   

7.
A multi-functional enzyme ICChI with chitinase/lysozyme/exochitinase activity from the latex of Ipomoea carnea subsp. fistulosa was purified to homogeneity using ammonium sulphate precipitation, hydrophobic interaction and size exclusion chromatography. The enzyme is glycosylated (14–15%), has a molecular mass of 34.94 kDa (MALDI–TOF) and an isoelectric point of pH 5.3. The enzyme is stable in pH range 5.0–9.0, 80 °C and the optimal activity is observed at pH 6.0 and 60 °C. Using p-nitrophenyl-N-acetyl-β-d-glucosaminide, the kinetic parameters Km, Vmax, Kcat and specificity constant of the enzyme were calculated as 0.5 mM, 2.5 × 10−8 mol min−1 μg enzyme−1, 29.0 s−1 and 58.0 mM−1 s−1 respectively. The extinction coefficient was estimated as 20.56 M−1 cm−1. The protein contains eight tryptophan, 20 tyrosine and six cysteine residues forming three disulfide bridges. The polyclonal antibodies raised and immunodiffusion suggests that the antigenic determinants of ICChI are unique. The first fifteen N-terminal residues G–E–I–A–I–Y–W–G–Q–N–G–G–E–G–S exhibited considerable similarity to other known chitinases. Owing to these unique properties the reported enzyme would find applications in agricultural, pharmaceutical, biomedical and biotechnological fields.  相似文献   

8.
Cladoptosis, the abscission of twigs, is the main mechanism of changes in crown structure in senescing pedunculate oak (Quercus robur L.). We tested the hypotheses that abscission zones in nodes of old pedunculate oak trees reduce leaf-specific hydraulic conductance of shoots and thereby limit the stomatal conductance and assimilation.Hydraulic conductance and leaf-specific hydraulic conductance, measured with a high pressure flowmeter in 0.5–1.5 m long shoots, were significantly lower in shoots of low vigour compared to vigorous growing shoots in a 165-years-old stand in the southeast of Germany. Two types of bottlenecks to water transport could be identified in shoots of old oak trees, namely nodes and abscission zones. In young twigs, vessel diameter and vessel density in nodes with abscission zones were significantly reduced compared with internodes. In nodes without abscission zones, vessel density was significantly reduced. The reduction of hydraulic conductance was especially severe in the smallest and youngest shoots with diameters less than 2 mm. Internodes of 1–5 mm sapwood diameter had an average hydraulic conductance of 7.13×10−6±0.2×10−6 kg s−1 m−1 MPa−1, compared to 4.54×10−6±0.3×10−6 kg s−1 m−1 MPa−1 in those with nodes.Maximum stomatal conductance and maximum net assimilation rate increased significantly with hydraulic conductance and leaf-specific hydraulic conductance. Maximum rate of net photosynthesis Amax of the most vigorous shoots (VC0) (7.34±0.55 μmol m−2 s−1) was significantly higher (P<0.001) than in shoots of other vigour classes (5.97±0.28 μmol m−2 s−1). Our data support the hypothesis that the changes in shoot and consequently crown architecture that are observed in ageing and declining trees can limit photosynthesis by reducing shoot hydraulic conductance. Abscission zones increase the hydraulic disadvantage of less vigorous compared to vigorously growing twigs. Cladoptosis might serve as a mechanism of selection between twigs of different efficiency.  相似文献   

9.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

10.
Biodegradation of Methyl red by Galactomyces geotrichum MTCC 1360   总被引:1,自引:0,他引:1  
Galactomyces geotrichum MTCC 1360 can decolorize triphenylmethane, azo and reactive high exhaust textile dyes. At shaking condition this strain showed 100% decolorization of a toxic azo dye Methyl red (100 m gl−1) within 1 h in deionized water at 30 °C. The degradation of Methyl red was possible through a broad pH (3–12) and temperature (5–50 °C) range. Glucose and mycelium concentration had increased the decolorization rate, but the addition of 1 gl−1 molasses in deionized water made decolorization possible in only 10 min. Induction in the NADH–dichloro phenol indophenol (NADH–DCIP) reductase, Malachite green reductase, laccase and lignin peroxidase (Lip) activities were observed in the cells obtained after complete decolorization, showing that there is direct involvement in the degradation of Methyl red. The absence of N-N′-dimethyl-p-phenylenediamine (DMPD) in 5 °C, 2-aminobenzoic acid (ABA) in 50 °C and both the compounds in 30 °C sample have shown the differences in the metabolic fate of Methyl red at different temperatures. The untreated dye at 300 mg l−1 concentration showed 88% germination inhibition in Sorghum bicolor, whereas it was 72% in Triticum aestivum. There was no germination inhibition for both the plants by Methyl red metabolites at 300 mg l−1 concentration.

The scientific relevance of the paper

The azo dye Methyl red (100 mg l−1) was decolorized by G. geotrichum MTCC 1360 within 1 h at shaking condition in deionized water. This organism could decolorize Methyl red at wide pH and temperature ranges. Decolorization time was reduced to 10 min by the addition of molasses to deionized water. There was induction in laccase and Lip, NADH–DCIP reductase and Malachite green reductase activities. The metabolic fate of Methyl red changes with temperature which can be evidenced by the formation of 2-ABA at 5 °C, N-N′-DMPD at 50 °C and both the compounds were absent at 30 °C. Phytotoxicity showed that metabolites of dye had induced shoot and root length of both the tested plants.  相似文献   

11.
A sensitive, selective, and rapid enzymatic method is proposed for the quantification of hydrogen peroxide (H2O2) using 3-methyl-2-benzothiazolinonehydrazone hydrochloride (MBTH) and 10,11-dihydro-5H-benz(b,f)azepine (DBZ) as chromogenic cosubstrates catalyzed by horseradish peroxidase (HRP) enzyme. MBTH traps free radical released during oxidation of H2O2 by HRP and gets oxidized to electrophilic cation, which couples with DBZ to give an intense blue-colored product with maximum absorbance at 620 nm. The linear response for H2O2 is found between 5 × 10−6 and 45 × 10−6 mol L−1 at pH 4.0 and a temperature of 25 °C. Catalytic efficiency and catalytic power of the commercial peroxidase were found to be 0.415 × 106 M−1 min−1 and 9.81 × 10−4 min−1, respectively. The catalytic constant (kcat) and specificity constant (kcat/Km) at saturated concentration of the cosubstrates were 163.2 min−1 and 4.156 × 106 L mol−1 min−1, respectively. This method can be incorporated into biochemical analysis where H2O2 undergoes catalytic oxidation by oxidase. Its applicability in the biological samples was tested for glucose quantification in human serum.  相似文献   

12.
The rate constant for the hydrolysis of prostacyclin (PGI2) to 6-keto-PGF was measured by monitoring the UV spectral change, over a pH range 6 to 10 at 25°C and the total ionic strength of 0.5 M. The first-order rate constant (kobs) extrapolated to zero buffer concentration follows an expression, kobs = kH+ (H+), where kH+ is a second-order rate constant for the specific acid catalyzed hydrolysis. The value of kH+ obtained (3.71 × 104 sec−1 M−1) is estimated approximately 700-fold greater than a kH+ value expected from the hydrolysis of other vinyl ethers. Such an unusually high reactivity of PGI2 even for a vinyl ether is attributed to a possible ring strain release that would occur upon the rate controlling protonation of C5. A Brønsted slope (α) of 0.71 was obtained for the acid (including H3O+) catalytic constants, from which a pH independent first-order rate constant for the spontaneous hydrolysis (catalyzed by H2O as a general acid) was estimated to be 1.3 × 10−6 sec−1. An apparent activation energy (Ea) of 11.85 Kcal/mole was obtained for the hydrolysis at pH 7.48, from which a half-life of PGI2 at 4°C was estimated to be approximately 14.5 min. when the total phosphate concentration is 0.165 M (cf. 3.5 min. at 25°C).  相似文献   

13.
The inversion of sucrose with β- -fructofuranosidase (EC 3.2.1.26) immobilized by an ionic bond on bead cellulose containing weak basic N,N-diethylamino-2-hydroxypropyl groups has been investigated. The immobilized enzyme is strongly bound at an ionic strength up to 0.1 M in the pH range 3–6. The amount adsorbed is proportional to porosity and to the exchange capacity of the ion exchange cellulose, reaching values up to 200 mg/g dry carrier, with an activity in 10% sucrose solution at 30°C, pH 5, >8000 μmol min−1 g−1. The inversion of sucrose with immobilized β- -fructofuranosidase was carried out in a stirred reactor. The dependence of activity on pH (3–7), temperature (0–70°C) and concentration of the substrate (2–64 wt%) were determined, and the inversion was compared with that obtained using non-immobilized enzyme under similar conditions. The rate of inversion at low substrate concentration (2–19 wt%) was described by Michaelis-Menten kinetics.  相似文献   

14.
Two processes of porphyrins in heterogeneous systems containing aqueous and membrane phases have been studied with hematoporphyrin and hematoporphyrin derivative: Dimerization equilibrium in the aqueous phases and porphyrin-membrane binding equilibrium using liposomes as models for biological membranes. The interrelationship of aqueous aggregations and membrane binding was probed and the porphyrin aggregation state in the membrane, at equilibrium, was assessed. Fluorimetric techniques were employed. The dimerization equilibrium constants, at neutral pH and 37°C were found to be 2.8 · 105 M−1 and 1.9 · 106 M−1 for hematoporphyrin and its derivative, respectively. Over a porphyrin concentration range going from monomer-dominant to dimer-dominant systems, we have found that only monomers are bound to the membrane. The respective monomer-liposome binding constants, found to be independent of the initial monomer/dimer distribution in the aqueous phase, were determined to be 1.6 · 103 M−1 and 4.1 · 103 M−1 at neutral pH and 37°C for hematoporphyrin and its derivative, respectively. The monomer-liposome interaction was found to perurb the initial monomer/dimer distribution in the aqueous phase, so that the monomers residing at equilibrium in the membrane originate from both monomers and dimers in the aqueous phase.  相似文献   

15.
A highly sensitive, kinetically unambiguous assay for α-factor-induced delay of cell passage through the “start” step of cell division in yeast is presented. The assay employs perfusion with periodic microscopy to monitor the bud emergence kinetics on the 20% of cells within an exponentially growing population which exist prior to the α-factor execution point of start. The t1/2 for cell passage through start by this population of cells is 31 min in the absence of α-factor. The inhibition constant, KI, represents the α-factor concentration which produces a 50% inhibition of this rate and is equal to 2×10−10M. A second assay for maximal cell division arrest by α-factor on whole populations of cells is presented. This assay shows a maximum cell division arrest time of 125±5 h at saturating α-factor, and a K50 (that is, an α-factor concentration which produces a half-maximal response) of 2.5×10−8M. Both assays were performed in the effective absence of α-factor inactivation. Values of the dissociation constant KD and total number of receptors per cell which specifically mediate cell division arrest or delay were estimated to be 2.5×10−8M and 104, respectively. These estimates, along with the quantitative dose-response data for division arrest which are presented here, are consistent with each receptor·α-factor complex which is present on the cell at equilibrium producing a 43±10 s delay of cell passage through start. Surprisingly, this number is constant within twofold over the entire range of cellular division arrest responses to α-factor, that is, from a 1.9-fold inhibition of the rate of cell passage through start at 0.17 nM α-factor to a 125±5 h maximum arrest at saturating α-factor concentrations of >170 nM. The possible significance of this observation toward the mechanism of α-factor-induced cell division arrest is discussed.  相似文献   

16.
Galactomannans isolated from legume seed endosperms, including those of commercial interest, have been characterized by multidetection aqueous SEC. Galactomannans derived from seeds of the Faboideae subfamily had substantially higher Mw than those from Caesalpinioideae seeds (Mw,Fab = 2.4–3.1 × 106 g/mol, Mw,Caes. = 0.86–2.1 × 106 g/mol) and within the latter botanical subfamily, an apparent correlation between Mw and the degree of galactose substitution DG was found. The molar mass distributions were unimodal and differed primarily by a scale factor, with distributional widths narrower than a true Flory ‘most-probable distribution’; good fits to Schulz–Zimm model were obtained. Across subfamilies no differences were found in the exponents of [η]–M and RvM relationships (0.61 ± 0.02, 0.54 ± 0.01, respectively), the Flory chain stiffness ratio (C = 20 ± 1 (BSF analysis)), or the persistence length (Lp = 5.5 ± 0.2 nm) obtained from SEC fraction data. However, it was found that prefactors in the [η]–M and RvM relationships as well as the unperturbed parameter KΘ decrease in proportion to DG and therefore chain density. Generalized relationships incorporating galactose-dependent prefactors were therefore developed to model SEC fraction data of native galactomannans ([η]GM = (1800 ± 200) × Mo−1.61 × M0.61±0.02, Rv,GM = 0.63 ± 0.05 × Mo−0.54 × M0.54±0.01) as well as lower-M fractions obtained by ultrasonication ([η]GM = (730 ± 100) × Mo−1.71 × Mw0.71±0.02, Rv,GM = 0.49 ± 0.05 × Mo−0.57 × Mw0.57±0.01, M ≈ 1 × 105-native). As a consequence of this dependence and the observed patterns in molar mass variation, [η] varies within a narrow range for galactomannans as a whole despite substantial Mw differences.  相似文献   

17.
You Wang  Xuexi Tang   《Harmful algae》2008,7(1):65-75
Interactions between Prorocentrum donghaiense Lu and Scrippsiella trochoidea (Stein) Loeblich III, two species of causative bloom dinoflagellates in China, were investigated using bi-algal cultures under controlled laboratory conditions. The growth of P. donghaiense and S. trochoidea were significantly suppressed when the initial cell densities were set at 1.9 × 104 cells mL−1 or 1.9 × 105 cells mL−1 for P. donghaiense and 1.0 × 104 cells mL−1 for S. trochoidea when the initial size/density ratio was 1:1 or 10:1, respectively, but no out-competement was observed in either bi-algal culture by the end. The simultaneous assay on the culture filtrate showed that P. donghaiense filtrate prepared at a lower initial density (1.9 × 104 cells mL−1) stimulated the co-cultured S. trochoidea at a density of 1.0 × 104 cells mL−1, but filtrate at a higher density (1.9 × 105 cells mL−1) depressed its growth. Differently, the filtrate of S. trochoidea at a density of 1.0 × 104 cells mL−1 significantly suppressed the growth of P. donghaiense at a density of 1.9 × 104 cells mL−1, but had little stimulatory effect on P. donghaiense at a density of 1.9 × 105 cells mL−1compared to the control (P > 0.05). It is likely that these two species of microalgae interact with each other mainly by releasing allelochemical substance(s) into the culture medium, and a direct cell-to-cell contact was not necessary for their mutual interaction. We then quantify their interactions in the bi-algal culture by using a mathematical model. The estimated parameters from the model showed that the inhibition exerted by S. trochoidea on P. donghaiense was about 43 and 24 times stronger than the inhibitory effect that P. donghaiense exerted on S. trochoidea when the initial size/density were 1:1 and 10:1, respectively. S. trochoidea seemed to have a survival strategy that was superior to P. donghaiense in the bi-algal culture under controlled laboratory conditions. We also observed a closely positive relationship between the initial cell density and its effect on the co-cultured microalga by measuring the fluorenscence: filtrate prepared from higher initial cell density had stronger interference on the co-cultured microalga. Moreover, pre-treated under different temperature conditions (30 °C, 60 °C and 100 °C) would significantly changed the effect of culture filtrate on the co-cultured microalga. Result inferred that P. donghaiense or S. trochoidea would release allelochemicals into the bi-algal culture medium and the allelochemicals might be a mixture with temperature-sensitive components in it.  相似文献   

18.
The kinetics of formation of the complex ion, μ-carbonato-di-μ-hydroxo-bis((1,5-diamino-3-aza-pentane) cobalt(III), from the tri-μ-hydroxo-bis((1,5-diamino-3-aza-pentane(III)cobalt(III)) ion in aqueous buffered carbonate solution have been studied spectrophotometrically at 295 nm over the ranges 20.0θ°C34.8, 8.03pH9.44, 5 mM [CO32−35 mM and at an ionic strength of 0.1 M (LiClO4). On the basis of the kinetic results a mechanism, involving rapid cleavage of an hydroxo bridge followed by carbon dioxide uptake with subsequent bridge formation, has been proposed. At 25 °C, the rate of the carbon dioxide uptake is 0.58 M−1 s−1 with ΔH≠ = (13.2±0.7) kcal mol−1 and ΔS≠ = (−15.1 ± 0.7) cal deg−1 mol−1. The results are composed with those obtained for several mononuclear cobalt(III) and one dinuclear cobalt(III) complexes.  相似文献   

19.
A novel raw starch degrading α-cyclodextrin glycosyltransferase (CGTase; E.C. 2.4.1.19), produced by Klebsiella pneumoniae AS-22, was purified to homogeneity by ultrafiltration, affinity and gel filtration chromatography. The specific cyclization activity of the pure enzyme preparation was 523 U/mg of protein. No hydrolysis activity was detected when soluble starch was used as the substrate. The molecular weight of the pure protein was estimated to be 75 kDa with SDS-PAGE and gel filtration. The isoelectric point of the pure enzyme was 7.3. The enzyme was most active in the pH range 5.5–9.0 whereas it was most stable in the pH range 6–9. The CGTase was most active in the temperature range 35–50°C. This CGTase is inherently temperature labile and rapidly loses activity above 30°C. However, presence of soluble starch and calcium chloride improved the temperature stability of the enzyme up to 40°C. In presence of 30% (v/v) glycerol, this enzyme was almost 100% stable at 30°C for a month. The Km and kcat values for the pure enzyme were 1.35 mg ml−1 and 249 μM mg−1 min−1, respectively, with soluble starch as the substrate. The enzyme predominantly produced α-cyclodextrin without addition of any complexing agents. The conditions employed for maximum α-cyclodextrin production were 100 g l−1 gelatinized soluble starch or 125 g l−1 raw wheat starch at an enzyme concentration of 10 U g−1 of starch. The α:β:γ-cyclodextrins were produced in the ratios of 81:12:7 and 89:9:2 from gelatinized soluble starch and raw wheat starch, respectively.  相似文献   

20.
Brucella abortus strain RB51 (SRB51) is a new cattle vaccine that is approved for use in the U.S. for prevention of brucellosis. At the present time, other countries are implementing or considering the use of SRB51 vaccine in their brucellosis control programs. In the current study, the effect of three stabilizing media, two fill volumes (1 and 3 ml), and three storage temperatures (−25, 4 and 25°C) on the viability of lyophilized SRB51 over a 52 week period was determined. The effects of three concentrations of bacteria (5×108, 1×109, or 5×109 cfu/ml) and two storage temperatures (4 or 25°C) on viability of liquid SRB51 vaccine were also determined. For lyophilized strain RB51 vaccine, fill volume did not influence viability (P> 0·05) during lyophilization. Although fill volume did not influence viability during storage in World Health Organization (WHO) media or media containing both WHO and Lactose Salt (LS) media, 1 ml fill volumes of SRB51 in LS media had greater (P< 0·05) viability when compared to 3 ml fill volumes. Lyophilized SRB51 vaccine stored at 25°C had a more rapid decline in viability (P< 0·05) when compared to vaccine stored at −25 or 4°C. With the exception of the 3-ml fill volumes of LS media, all three stabilizing media were similar in maintaining viability of SRB51 at −25°C storage temperatures. However, when compared to WHO or WHO/LS media, stabilization in LS media was associated with a more rapid decline in viability during storage at 4 or 25°C (P< 0·05). Initial SRB51 concentration in liquid vaccine did not influence (P> 0·05) viability during storage at 4 or 25°C. When compared to liquid SRB51 vaccine stored at 25°C, storage at 4°C was associated with a slower decline in viability (P< 0·05) during 12 weeks of storage. Biochemical and morphological characteristics of SRB51 were stable under the storage conditions utilized in the present study. This study suggests that viability of SRB51 can be readily maintained during storage as a lyophilized or liquid brucellosis vaccine.  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