首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Yeast alcohol dehydrogenase (ADH) solutions (approximately 1 mg/ml, pH 7) were sheared in a coaxial cylindrical viscometer. This was fitted with a lid sealing the contents from the atmosphere and preventing evaporation. At 30 degrees C after a total of 5 hr intermittent shearing at 683 sec-1 no losses of activity were observed. No losses were found after 5 hr continuous shearing and in a no-shear control. At 40 degrees C and 683 sec-1 there were only small activity losses in 5 hr. Shearing at 3440 sec-1 no measurable losses of activity were found with a 1.03 mg/ml solution in 5 hr at 30 degrees C, a 1.03 mg/ml solution in 8 hr at 5 degrees C, and with a 3.89 mg/ml solution in 3 hr at 5 degrees C. In all these cases, however, a white precipitate formed that was not observed in zero shear control experiments. The sheared 3.89 mg/ml solution was clarified by centrifugation. It was shown that there were no ADH aggregates in the supernatant and that the precipitate was less than 2% of the original protein. At 30 degrees C under adverse pH conditions (pH 8.8) there was no significant difference in activity losses of an approximately 1 mg/ml solution sheared at 65 and 744 sec-1. An approximately 0.5 mg/ml ADH solution, pH 7, was agitated in a small reactor with no free air-liquid interface. Peak shear rates near the impeller were estimated to be about 9000 sec-1. Only a small decrease in specific activity was observed until over 15 hr total running at 5 degrees C.  相似文献   

2.
The acid-catalyzed hydrolysis of heparin from Cu(II) complex was studied as a function of time and temperature. Four independent calculations showed that the hydrolysis, during the 5-hr period examined, obeys the first-order kinetic law. Specific rate constants, calculated at 50°C, 57°C, 65°C, 71°C, and 80°C, were 3.3 × 10?5 sec?1, 6.5 × 10?5 sec?1, 10.4 × 10?5 sec?1, 15.1 × 10?5 sec?1, and 26.6 × 10?5 sec?1, respectively. Arrhenius plots of the data yielded 14.7 kcal as the energy of activation. An independent run of the self-hydrolysis of heparin at 57°C also obeyed first-order kinetics and its specific rate constant of 6.4 × 10?5 sec?1 is in excellent agreement with that of the hydrolysis of Cu(II)-heparin at 57°C. The anticoagulant activity of heparin and of the Cu(II)-heparin are not appreciably different. Further, the inactivation of heparin closely parallels Cu(II) release from the Cu(II) complex which in turn parallels desulfation.  相似文献   

3.
4.
The viscosity change of myosin A concentrated solution with or without other components was measured as the incubation time elapsed at 30°C.

The viscosity of myosin A solution increased, but that of F-actin solution did not. The shear stress at 0.04 sec?1 was not increased to 1.0 dyne/cm2 in the former, but in the latter was below 0.5 dyne/cm2.

The viscosity of myosin B solution increased slightly, but that of native tropomyosin-free myosin B solution decreased remarkably. In both the shear stress at 0.04 sec?1 was greater than or equal to 15 dynes/cm2.

The speed of the viscosity increase in the presence of 3 mm pyrophosphate and 3 mm MgCl2 was higher in concentrated solution of myosin B than in that of native tropomysin-free myosin B. The shear stress at 0.04 sec?1 after 6 hr at 30°C was 11.5 and 8.2 dynes/cm2, respectively.

The effect of native tropomyosin and actin on the viscosity change was discussed.  相似文献   

5.
The specific activity (k′1) and concentration of red blood cell catalase from four inbred strains of mice (BALB/c, C57BL, C57BL/6, and NBL) were measured to determine the mechanisms responsible for interstrain variations in enzyme activity. The specific activities of RBC catalase in NBL and the C57BL sublines are equal (2.5×107 m ?1 sec?1), while that of BALB/c (4.0×107 m ?1 sec?1) is 67% greater. The relative concentration of catalase is approximately 30% lower in NBL erythrocytes compared to the other three strains. The activity of BALB/c RBC catalase is due to a high k′1 coupled with a high intracellular concentration; RBC catalase activity in the C57BL sublines is the result of a low k′1 and high concentration. A low k′1 and a low concentration are responsible for the low catalase activity levels found in NBL erythrocytes.  相似文献   

6.
Genes of β-mannosidase 97 kDa, GH family 2 (bMann9), β-mannanase 48 kDa, GH family 5 (bMan2), and α-galactosidase 60 kDa, GH family 27 (aGal1) encoding galactomannan-degrading glycoside hydrolases of Myceliophthora thermophila C1 were successfully cloned, and the recombinant enzymes were purified to homogeneity and characterized. bMann9 displays only exo-mannosidase activity, the K m and k cat values are 0.4 mM and 15 sec?1 for p-nitrophenyl-β-D-mannopyranoside, and the optimal pH and temperature are 5.3 and 40°C, respectively. bMann2 is active towards galac-tomannans (GM) of various structures. The K m and k cat values are 1.3 mg/ml and 67 sec?1 for GM carob, and the optimal pH and temperature are 5.2 and 69°C, respectively. aGal1 is active towards p-nitrophenyl-α-D-galactopyranoside (PNPG) as well as GM of various structures. The K m and k cat values are 0.08 mM and 35 sec?1 for PNPG, and the optimal pH and temperature are 5.0 and 60°C, respectively.  相似文献   

7.
The molecular parameters of pronase-treated acid-soluble bovine skin collagen (P-ASC) were determined from low-shear gradient viscosity, electric birefringence, and electron microscopic data in order to determine the shear gradient range in which viscosity studies yield data which can be correctly interpreted by use of the various hydrodynamic equations for prolate ellipsoids of revolution. The P-ASC solutions could be characterized by a single relaxation process in electric briefringence with rotary diffusion coefficient θ of 810 sec?1 and a corresponding molecular length of 2850 Å. Viscosity data were found to be shear gradient dependent and only the extrapolated zero-shear value [η]D = 0 could be used with the viscosity hydrodynamic equations to provide a correct value of molecular length. Intrinsic viscosities obtained at shear gradients >250 sec?1 are nearly 30% lower than the zero-shear value. Untreated acid-soluble collagen (ASC) solutions contain aggregates and these appear, from electric birefringence data, to be of endlinked character. ASC solutions show a much more marked shear gradient dependence than P-ASC. For example, at D~500sec?1,[η] = 22 dl/g, whereas the extrapolated zero-shear value of[η] was found to be 44 dl/g. Thus, the shear gradient dependence of native collagen solutions is much more marked than previously assumed and, in contrast to the usual practice, only viscosities measured near zero shear can be interpreted in terms of molecular parameters for collagen solutions containing aggregates.  相似文献   

8.
Cotton fabric was first oxidized with sodium periodate, and then employed to immobilize catalase. Optimization studies for oxidation of the fabric and immobilization of the enzyme were performed. The properties of the immobilized catalase were examined and compared with those of the free enzyme. A high activity of the immobilized enzyme was obtained when the fabric was oxidized at 40°C and pH 6.0 for 8h in a bath containing 0.20 mol L?1 sodium periodate and the enzyme was immobilized at 4°C for 24h with a catalase dosage of 120.0 U mL?1. The immobilized enzyme exhibited optimum activity at 40°C, while the free enzyme had optimal temperature of 30°C, suggesting that the immobilized catalase could be used in a broader temperature range. Both the immobilized and free enzyme had pH optima of 7.0. The staining test and reusability showed that the catalase was fixed covalently on the oxidized cotton fabric.  相似文献   

9.
The translational diffusion coefficient of CF1 at low and high protein concentration as well as at different ionic strength (0.05 – 1.65 M) wsa determined by means of quasi-elastic light scattering experiments. The diffusion coefficient changes from D20,wo = 3.12 × 10?7 cm2 · sec?1 at 0.05 M, pH 7.8, 20°C, to D20,wo = 3.52 × 10?7 cm2 · sec?1 at 1.6 M, pH 7.8, 20°C. At high enzyme concentration (20 mg/ml) and under crystallization conditions (Paradies, BBRC 91: 685, 1979) CF1 behaves as a solution of “true” hard spheres, whereas at low salt concentration the ionic atmosphere has a larger spatial extent, resulting in a higher effective hydrodynamic radius (RH = 65 Å).  相似文献   

10.
Flow dichroism of DNA: a new apparatus and further studies   总被引:3,自引:0,他引:3  
P R Callis 《Biopolymers》1969,7(3):335-352
A new apparatus for the study of flow dichroism of macromolecules is described. The flow is down a long, narrow channel and an unpolarized light beam propagates along the flow direction. For a molecule such as DNA, in which the transition moments of the chromophores are perpendicular to the axis of orientation, an increase of absorbance is observed during flow. The apparatus is best suited for macromolecules which are readily orientable or at high shear gradients so that the extinction angle is close to 0°. The apparatus has the following advantages: dilute macromolecule solutions can be used; high shear gradients are easily obtained; only small volumes of solution are needed. The flow can be stopped rapidly so that relaxation times for disorientation can be studied. The flow dichorism of native, two-stranded DNA has been measured for the molecular weight range of 0.6 × 106 to 125 × 106, and for the shear gradient range (in aqueous solution at 25°C) from 200 sec?1 to 21000 sec?1. At a fixed gradient the dichroism increases with molecular weight, but the curve is concave downwards. At a given molecular weight the dichroism increases with increasing shear gradient, but the curve is concave downwards. When the solvent viscosity and temperature are varied, the dichroism is a function of η〈G〉/T showing that the orientation is due to hydro-dynamic shear stress and that the flexibility of DNA in a flow field is not due to local denaturation. The Zimm-Rouse theory with no parameters taken from flow optical data predicts the correct order of magnitude of the dichroism but the experimentally observed shear gradient and molecular weight dependence do not fit the theory. This is an expected result, since the theory is believed to be applicable only at small distortions and extensions of the macromolecule.  相似文献   

11.
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°C, and was stable between pH 4.5-10.0 and under 50°C. The K m and V max were measured with a chitosan of a D.A. of 20.2% as 0.52 mg/ml and 7.71×10?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, k 1 values, of 4.98×10?4, 2.3×10?4, and 9.3×10?6 sec?1, respectively. The enzyme hardly catalyzed degradation of chitooligomers smaller than the pentamer.  相似文献   

12.
The kinetics of formation of the intermediate complex between catalase and H2O2 has been reexamined. It has been shown that the kinetics consists of a rapid and of a subsequent slow phase. At the maximum of the transient decrement of the optical absorption, the system was found to be in a terminal state with regard to the rapid phase. On this basis, the formation curve of the intermediate complex was calculated. From the parameters of the curve the maximal saturation of catalase hematins (from horse erythrocytes) by H2O2 is 35%. The absolute spectrum of the intermediate complex was established. The variation of the previously calculated rate constant of formation of the intermediate complex was shown to be due to the inapplicability of the pre-steady-state approximation to the rate data. By applying a more general approach and by the use of a computer, the individual rate constants of the peroxidatic scheme were calculated (relevant to micromolar solutions of catalase) k1 = (3.0 ± 0.2) × 106 M?1 sec?1k4 = (5.6 ± 0.3) × 106 M?1 sec?1 These values are 2.2 times higher in a nanomolar solution.  相似文献   

13.
The reaction of cytochalasin A with sulfhydryl groups was examined. Cysteine and glutathione reacted readily with cytochalasin A at pH 7.0, 20°C, following second-order kinetics with rate constants of 7,600 M?1 sec?1 and 870 × 103 M?1 sec?1. No reaction of cytochalasin B could be demonstrated under the same conditions. The reaction of cytochalasin A with the amino group of glycine ethyl ester had a second-order rate constant of 0.02 M?1 sec?1. Cytochalasin A did not react with sufhydryl groups of native ovalbumin or lactic dehydrogenase but reacted with an least 2 and 12 groups respectively when the proteins were denatured in 0.1% SDS. The reactivity of cytochalasin A with sulfhydryl groups is attributable to the α,β-unsaturated ketone groups it contains.  相似文献   

14.
The hydrolyses of p-nitrotrifluoroacetanilide catalyzed by water and imidazole were examined at 70°C. The pH-rate constant profile of the hydrolysis in H2O was examined in the pH range 0.0–11.4. The hydrolysis was independent of pH in the region from pH 1.0 to 4.5, presumably a water-catalyzed reaction. The rate constant and the D2O solvent isotope effect for this reaction were 1.0 × 10?4 sec?1 and 3.7, respectively. Both natural imidazole and imidazolium cation catalyzed hydrolysis. The rate constant of the hydrolysis catalyzed by neutral imidazole was determined to be 5.4 × 10?3M?1 sec?1 and the D2O solvent isotope effect was 1.8.  相似文献   

15.
The objective of this study was to analyze the mechanism of some physiological processes accompanying acquisition of sunflower (Helianthus annuus L.) chilling resistance due to seeds hydropriming in the presence of salicylic acid, jasmonic acid, 24-epibrassinolide followed exposition of seeds to short-term heat shock treatment. The seeds were hydroprimed at 25 °C in limited amounts of water or solution of salicylic or jasmonic acid at 10?2, 10?3 and 10?4 M concentration, 24-epibrassinolide at 10?6, 10?8 and 10?10 M concentration. The seeds were incubated for 2 days, subjected to short-term heat shock (45 °C, 2 h) and chilled for 21 days at 0 °C. Sunflower chilling susceptibility and physiological responses were evaluated according to the inhibition of radicle growth, the inhibition of the number of lateral roots formation, the activity of catalase and changes in soluble carbohydrates in seedlings developing for 72 h at 25 °C. Hydropriming and short-term heat shock application explicitly reduced inhibition of roots as well as lateral roots development by allowing the germinating seeds to recover from the growth-inhibiting effects of chilling. Seeds hydropriming in solutions containing salicylic acid, jasmonic acid and 24-epibrassinolide followed heat shock treatment additionally promoted the activity of catalase and sugars metabolism, which stimulated seedlings development and alleviated the decrease of F v/F m caused by chilling conditions. These beneficial effects contributed to increased resistance of sunflower seedlings to chilling stress. The present study demonstrated that the most profitable effect on reducing negative effect of chilling may be achieved by short-term heat shock applied during hydropriming in water supplemented with 24-epiBL (10?8 and 10?10 M) or salicylic acid (10?3 and 10?4 M).  相似文献   

16.
The effects of alkalic salts on the apparent viscosity of acid precipitated protein (APP) of soybean-suspending systems and heated gels were investigated using a modified coaxial cylinder viscometer. A hybrid program was established. (i) a cyclic temperature test (20 90 20°C) under a constant shear rate and (ii) a cyclic shearing test (48.7 243.7 48.7 sec–1) under isothermal conditions. The apparent viscosity of protein suspending systems (12%, wt/vol) gradually decreased with increasing temperature to about 70°C. The apparent viscosity increased with a rise in temperature in the range of 70 to 90°C, in ascending order of Hofmeister’s series.

With a fall in temperature, the apparent viscosity increased considerably in this order. The effect of anions to apparent viscosity was larger than that of cations at all measured temperatures in the gel-forming period.  相似文献   

17.
The present study describes a strain of Gloeocapsa sp. designated as Gacheva 2007/R‐06/1, originally isolated from a geothermal flow located in Rupite, Bulgaria. To evaluate whether this cyanobacterium is locally adapted to hot environment or has the ability to tolerate lower temperatures, its growth, biochemical composition, enzyme isoforms and activity of the main antioxidant enzymes and proteases were characterized under various temperatures and two irradiance levels. The strain was able to grow over the whole temperature range (15–40°C) under two different photon fluence densities – 132 μmol photons m?2 s?1 (unilateral, low light, LL) and 2 × 132 μmol photons m?2 s?1 (bilateral, high light, HL). The best growth occurred at either 34°C and LL or at 36°C and HL, but significant growth inhibition was noted at 15°C and 40°C. Low temperature treatment (15°C) resulted in higher levels of total protein and an increased activity of manganese superoxide dismutase (MnSOD) and glutathione reductase, as compared to optimum growth temperatures. After simultaneous exposure to 15°C and HL, increases in lipid content and activity of iron superoxide dismutase and catalase (CAT) were also observed. Cultivation of cells at 40°C enhanced MnSOD, CAT and peroxidase activities, regardless of irradiance level. Increased total protein content and protease activity at 40°C was only associated with the HL treatment. Overall, these results indicate that Gloeocapsa sp. strain Gacheva 2007/R‐06/1 used different strategies to enable cells to efficiently acclimate and withstand adverse low or high temperatures. This strain obviously tolerates a wide range of temperatures below its natural habitat temperature, and does not seem to be locally adapted to its original thermal regime. It behaved as a thermotolerant rather than a thermophilic cyanobacterium, which suggests its wider distribution in nature.  相似文献   

18.
Effects of deuteration on the Raman spectrum of a tryptophan residue have been examined. The 1386 cm?1 line of deuterated tryptophan residue has been found to be useful for tracing the hydrogen-deuterium exchange reaction of this residue in a protein. An examination on bovine α-lactalbumin at pH 6.4 and at 20°C indicates that two of the four tryptophan residues exchange with a rate constant much greater than 9 × 10?4 sec?1, while the other two exchange with a rate constant of 4 × 10?5 sec?1. The latter two have been assigned to Trp 28 and Trp 108 of this protein. The kinetics of hydrogen-deuterium exchange reaction of completely “free” tryptophan residue have been examined by a proton magnetic resonance study on tryptophan itself. By taking the result of this examination into account, the chance of exposure to the solvent for Trp 28 or Trp 108 has been estimated to be 3 × 10?6 at pH 6.4 and at 20°C.  相似文献   

19.
G. R. Palmer  O. G. Fritz 《Biopolymers》1979,18(7):1659-1672
The shape of fibrin intermediate polymers as well as the rate of fibrinogen polymerization was studied using diffusion measured by quasielastic light scattering. After their length distribution was narrowed by gel filtration, the polymers yielded translational and rotational diffusion coefficients of 0.37 ± 0.05 × 10?7 cm2 sec?1 and 142 ± 32 sec?1, respectively. Theoretical considerations indicated the polymers to be rigid rods. The rate of polymerization of fibrinogen monitored by diffusion paralleled that provided by simulataneous intensity measurements. Both monitors indicated polymerization occurs most rapidly at 30°C.  相似文献   

20.
Extracellular catalases produced by fungi of the genusPenicillium, i.e.,P. piceum, P. varians, andP. kapuscinskii, were purified by consecutive filtration of culture liquids. The maximum reaction rate of H2O2 decomposition, the Michaelis constants, and specific catalytic activities of isolated catalases were determined. The operational stability was characterized by the effective rate of catalase inactivation during enzymatic reaction (k in at 30°C). The thermal stability was determined by the rate of enzyme thermal inactivation at 45°C (k in * , s-1). Catalase fromP. piceum displayed the maximum activity, which was higher than the activity of catalase from bovine liver. The operational stability of catalase fromP. piceum was twofold to threefold higher than the stability of catalase from bovine liver. The physicochemical characteristics of catalases of fungi are better than the characteristics of catalase from bovine liver and intracellular catalase of yeastC. boidinii.  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