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 15 毫秒
1.
Light scattering and viscometric studies have been carried out on two preparations, A and B, of rooster comb hyaluronate. Sedimentation rate studies have also been performed with A. Light scattering measurements in 0.2 m KCl for preparation A gave a molecular weight of 3.3 × 106 and for B, 1.0 × 106. In (0.1–0.3) M NaCl similar measurements gave a particle weight for A of (4.4–6.4 × 106 and for B (1.7–2.8 × 106. In 0.066 m CaCl2 molecular weight values of 9.5 × 106 for A and 1.7 × 106 for B were obtained. Thus in the presence of Na+ and Ca2+ ions aggregates of chains persisted into dilute solution. Measurements by light scattering on A and B in 4 m guanidinium chloride gave values in the same range as those obtained in 0.2 m KCl. Sedimentation rate studies on A gave values of 10.3 Svedbergs in 0.2 m KCl and 12.2 Svedbergs in 0.2 m NaCl and 0.066m CaCl2. The shear dependence of the viscosity was studied using a conicylindrical viscometer at shear rates between 0.5 and 20 s?1. Preparation A in 0.2 m KCl and NaCl yielded values for (νsp/cc→0 of 5000 and 7100 ml g?1 respectively in keeping with the tendency to aggregate. The behaviour for preparation B was similar. In 0.066 m CaCl2 there was a marked dependence of viscosity on shear speed below 10 s?1 for all concentrations and the value of (νsp/c)→0 at 0 s?1 for preparation A was 7700 ml g?1 while at a shear rate of 8 s?1 (νsp/c)c→0 ? 5000 ml g ?1. Similar effects were found for preparation B and the data suggest associations of chains disruptable by weak shear forces. The increase in viscosity with concentration in the presence of 0.066 m CaCl2 was much less than in the presence of KCl or NaCl, suggesting that the Ca2+ had a marked effect on the ”rigidity’ of the molecules in solution. A viscometric titration experiment with Ca2? showed that a level of 0.02 m CaCl2 in 0.2 m NaCl was sufficient to produce the change in viscosity presented above and that significant perturbations of the viscosity were present at 0.005?0.01 m CaCl2.  相似文献   

2.
1. The process of denaturation of the chicken muscle dimeric enzyme triosephosphate isomerase on addition of guanidinium chloride has been studied at pH 7.6, the pH at which the recovery of activity is optimal (100%) on removal of denaturant. Determinations of the sedimentation coefficient, intrinsic viscosity, molecular weight (by sedimentation equilibrium studies) and the absorption coefficient at 280 nm in various concentrations of guanidinium chloride concurred in showing a single, sharp transition at about 0.7 M guanidinium chloride at a protein concentration 1-5 mg/ml from the native enzyme to the dissociated, unfolded chains of the monomer. Relative fluorescent intensity measurements revealed a single transition at about 0.4 M guanidinium chloride at enzyme concentrations of about 0.05 mg/ml. 2. The process of denaturation in different guanidinium chloride concentrations was first order with respect to enzyme and about sixth order with respect to denaturant. 3. The rate of attainment of equilibrium during the renaturation obeyed second-order/first-order reversible kinetics. It was concluded that the rate-determining step in renaturation at pH 7.6 must be the association of two subunits.  相似文献   

3.
The myosin molecule was extracted from the smooth muscle parts of horse esophagus and purified by ammonium sulfate fractionation. The schlieren pattern of the sedimentation velocity run showed a very sharp single peak of.5.9. S (s20,w). Molecular weight of the protein was measured by means of the Archibald and sedimentation equilibrium methods, both in 0.5M KCI buffered by 1/150 M phosphate at pH 7.5 and at 5°C. The values obtained were 6.25 × 105 and 5.81 × 105respectively, for the two methods. The second virial coefficients were 1.1 × 104 and 1.2 × 10?4 ml/g. Denatured smooth muscle myosin was prepared in a solution of 5M guanidine HC1 containing 0.4 M KC1 and 0.2 M β-mercaptoet hanol buffered at pH 8.0. The weight-average molecular weight of the denatured smooth muscle myosin was 2.24 × 105 and the second virial coefficient was 7.6 × 10?4 ml/g. The values described above are in good agreement with those reported for rabbit skeletal myosin with ammonium sulfate fractionation. The molecular dimension of the molecule is estimated as the value for an axial ratio of 100, assuming a rigid rod molecular model for this molecule, both the thermodynamical and hydrodynamical treatment being in a good agreement with this estimation.  相似文献   

4.
The denaturation of immunoglobulin G and its light chains by guanidinium chloride at 25°C was followed using the dilatometric method. From results of the dilatometric measurements the differences between the partial molar volumes of the proteins in water and in guanidinium chloride solutions of various concentrations, respectively, have been obtained. The differences reflect the extent of unfolding as well as the denaturant binding to the protein. Several experiments were also performed in which the protein disulfide bonds were reduced by dithioerythritol.The volume change produced by the reduction of one mole of disulfide bonds of immunoglobulin G in 6 M guanidinium chloride was found to be the same as that for oxidized glutathione; the value was ?0.9 ml/mol.  相似文献   

5.
Wang X  Zhang X  Xu X  Zhang L 《Biopolymers》2012,97(10):840-845
Lentinan (β‐(1→3)‐D ‐glucan) was found to be successfully fractionated by the mixture of dimethyl sulfoxide (DMSO) and lithium chloride (LiCl) as a solvent and acetone as a precipitant. Light scattering and viscosity measurements were made on solutions of fractionated samples in pure DMSO and 0.2M LiCl/DMSO in the range of the molecular weight Mw from 21.7 × 104 to 84.7 × 104. The values of Mw in both solvents were almost the same, but the remarkable difference between the values of intrinsic viscosity [η] demonstrated that the LiCl/DMSO solvent greatly enhances the stiffness of the lentinan backbone. The observed intrinsic viscosity [η] was analyzed by the Yoshizaki‐Nitta‐Yamakawa theory of a worm‐like chain, and the persistence length q and molecular weight per unit contour length ML were determined roughly as 6.0 nm and 890 g nm?1 in 0.2M LiCl/DMSO, and 5.1 nm and 890 g nm?1 in pure DMSO, respectively. This slightly larger persistent length in 0.2M LiCl/DMSO also confirmed the higher stiffness of lentinan enhanced by the LiCl/DMSO solvent. The enhancement of the chain stiffness was ascribed to the electrostatic repulsion because of the hydrogen bonding of the hydroxyl protons of lentinan with the chloride ion, which is in turn associated with the Li+(DMSO)n macrocation complex. © 2012 Wiley Periodicals, Inc. Biopolymers 97: 840–845, 2012.  相似文献   

6.
The molecular weight of human serum low density lipoprotein (LDL) was determined for the first time by sedimentation equilibrium or, more accurately, flotation equilibrium, in high concentration of KBrNaBr containing Tris-Cl buffer plus EDTA (density = 1.20 – 1.49 g/ml). Assuming both the molecular weight and the partial specific volume were unknown, the results at different densities gave a value of 2.87 ± 0.12 × 106 for the molecular weight and 0.965 ± 0.014 ml/g for the partial specific volume.  相似文献   

7.
Eburnetoxin, a powerful vasoactive protein has been isolated from the venom of the marine snail Conus eburneus, monitored by the contractile effect to the rabbit aorta. The molecular weight was estimated to be 28, 000 by gel permeation chromatography and slab gel electrophoresis. The purified protein was electrophoretically homogeneous. The toxin at concentrations above 3 × 10?7 g/ml elicited a marked contractile response of aorta, which was inhibited by verapamil (10?6 M). The minimum lethal dose in the fish Rhodeus ocellatus smithi was 1 μg/g body weight.  相似文献   

8.
Hydrodynamic data, i.e. intrinsic viscosity and sedimentation coefficient, for denatured globular proteins in 6 M guanidinium chloride have been re-analysed in terms of the Yamakawa-Fujii theory of the wormlike cylinder model. Molecular parameters thus obtained ((mean value of R2 0/M) infinity, the Kuhn statistical segment length and the molecular weight per unit contour length) are in better agreement with the values obtained theoretically or by other methods than those evaluated on the basis of the model of non-draining random coil.  相似文献   

9.
Xylanases from alkalophilic thermophilic Bacillus spp. Wl and W2 were purified and characterized. The xylanases from the two strains were fractionated into two active components (I and II) by DEAE-Toyopearl 650M chromatography. Components I from the two strains had similar properties: optimum pH, 6.0; optimum temperature, 65°C; isoelectric point, pH 8.5 and 8.3; molecular weight, 21,500 and 22,500; and Michaelis constant, 4.5 and 4.0mg-xylan/ml. Components II from the two strains also had similar properties: optimum pH, 7.0~9.0 and 7.0~9.5; optimum temperature, 70°C; isoelectric point, pH 3.6 and 3.7; molecular weight, 49,500 and 50,000; and Michaelis constant, 0.95 and 0.57mg-xylan/ml. The activities of components I and II were inhibited by Hg++ and Cu++. Components I hydrolyzed xylan to yield xylobiose and higher oligomers, but components II produced xylose other than xylobiose and xylooligomers.  相似文献   

10.
DNA from untreated L-cells had a weight average molecular weight (Mw) of 5.7 ± 0.58·108 daltons as measured by sedimentation in an alkaline sucrose gradient. This value was reduced by one half after the cells were treated for 1 h with 8 μg/ml of N-methyl-N-nitrosourea (MNUA), 34 μg/ml of methyl methanesulfonate (MMS) or 0.16 μg/ml of N-methyl-N′-nitro-N-nitrosoguanidine (MNNG). That dose of MNUA produced 52 methylations per 5.7·108 daltons DNA. 20% of these were not purine derivatives and were assumed to contain some phosphotriesters. That dose of MMS (above) produced 290 methylations per 5.7·108 daltons DNA and about 14% of these were not purine derivatives. The rates of loss of methylated purines from DNA were 2.3% per hour for 7-methylguanine (7-MeG), 7.4% per hour for 3-methyladenine (3-MeA) and no detectable loss of O6-methylguanine (O6-MeG) over a 12 h period. Since phosphotriesters are alkali-labile the single-strand breaks probably arose from this structure and did not form within the cell. This conclusion is supported by the following considerations. MNUA was more effective than MMS at reducing the molecular weight of DNA, as measured in alkaline medium. The greater SN1 character of MNUA would cause a greater formation of phosphotriesters than would MMS.  相似文献   

11.
A novel agro-residue, tea stalks, was tested for the production of tannase under solid-state fermentation (SSF) using Aspergillus niger JMU-TS528. Maximum yield of tannase was obtained when SSF was carried out at 28 °C, pH 6.0, liquid-to-solid ratio (v/w) 1.8, inoculum size 2 ml (1?×?108 spores/ml), 5 % (w/v) ammonium chloride as nitrogen source and 5 % (w/v) lactose as additional carbon source. Under optimum conditions, tannase production reached 62 U/g dry substrate after 96 h of fermentation. Results from the study are promising for the economic utilization and value addition of tea stalks.  相似文献   

12.
Solution properties and structure of brain proteolipids   总被引:1,自引:0,他引:1  
R Zand 《Biopolymers》1968,6(7):939-953
Bovine white matter proteolipids have been studied by several physical methods and have been found to exist as micelles in 2 : 1 (v/v) chloroform–methanol solution. The data would indicate the existence of a critical micelle concentration at 0.017–0.022 g/100 ml. The curve appears linear in the range 0.017–0.2 g/100 ml, but from the data at higher concentrations it would appear that a change in slope is occurring in the region 0.2–0.3 g/100 ml. Light-scattering measurements on 2 : 1 (v/v) chloroform–methanol solutions containing more than 0.2 g/100 ml of proteolipid yielded a weight-average aggregate weight of 2.9 × 106and a radius of gyration of 64.5 Å. The intrinsic viscosity of the solutions was 0.32 dl/g and the Huggins constant was 1.085. Light-scattering measurements in 88.5% formic acid–0.5M sodium formate yielded a weight-average aggregate weight of 7.1 × 106 and a radius of gyration of 241 Å. The intrinsic viscosity observed for this solvent system is 0.14 dl/g and the Huggins constant is 1.005. Osmotic pressure measurements in 2 : 1 (v/v) chloroform–methanol containing less than 0.2 g/ 100 ml of proteolipid yielded a number-average aggregate weight of 7.2 × 104 Ultracentrifugal analysis in 1.5:1 (v/v) methylene chloride–methanol showed two broad peaks with, s values of s1.5% = 20.05 S, s2% = 19.79 S for the minor peak and s1.5%,2% = 1.86 S for the major peak. Optical rotatory dispersion studies revealed large changes in b0 with change in solvent and proteolipid concentration. The present data suggest that the mode of attachment of protein to lipid is primarily of a noncovalent type. The results of this investigation also suggest that the proteolipid micelle above 0.2 g/100 ml is cylindrical (prolate ellipsoid) in 2:1 (v/v) chloroform-methanol and approaches a more spherical shape in 88.5% formic acid. A structure for the proteolipid micellar complex above concentrations of 0.2 g/100 ml is proposed.  相似文献   

13.
More than 30 years ago, Nozaki and Tanford reported that the pK values for several amino acids and simple substances in 6 M guanidinium chloride differed little from the corresponding values in low salt (Nozaki, Y., and C. Tanford. 1967. J. Am. Chem. Soc. 89:736-742). This puzzling and counter-intuitive result hinders attempts to understand and predict the proton uptake/release behavior of proteins in guanidinium chloride solutions, behavior which may determine whether the DeltaG(N-D) values obtained from guanidinium chloride-induced denaturation data can actually be interpreted as the Gibbs energy difference between the native and denatured states (Bolen, D. W., and M. Yang. 2000. Biochemistry. 39:15208-15216). We show in this work that the Nozaki-Tanford result can be traced back to the fact that glass-electrode pH meter readings in water/guanidinium chloride do not equal true pH values. We determine the correction factors required to convert pH meter readings in water/guanidinium chloride into true pH values and show that, when these corrections are applied, the effect of guanidinium chloride on the pK values of simple substances is found to be significant and similar to that of NaCl. The results reported here allow us to propose plausible guanidinium chloride concentration dependencies for the pK values of carboxylic acids in proteins and, on their basis, to reproduce qualitatively the proton uptake/release behavior for the native and denatured states of several proteins (ribonuclease A, alpha-chymotrypsin, staphylococcal nuclease) in guanidinium chloride solutions. Finally, the implications of the pH correction for the experimental characterization of protein folding energetics are briefly discussed.  相似文献   

14.
Purification and characterization of the amylase of B. subtilis NRRL B3411   总被引:4,自引:0,他引:4  
The amylase of Bacillus subtilis NRRL B3411 has been purified and partially characterized. The specific activity can be increased from 300,000 units/g to 6,000,000 units/g with a 60% recovery of total units. The purified material consists of one major and one trace anodic component as determined by disc gel electrophoresis. The molecular weight was 48,000 as determined by bio-gel filtration; the molecular weight was 44,900 ± 2400 as determined by sedimentation equilibrium methods. This purified enzyme is stable at, 70°C in the presence of 0.01 M Ca++ and 0.1 M NaCl over a broad pH range from 5.5–9.5. The pH activity profile indicates optimum activity at pH 6.0. This amylase exhibits maximum activity at 60°C. The enzyme is a liquefying α-amylase as determined by analysis of hydrolysis products and immunological studies.  相似文献   

15.
Enolases (2-phospho-d-glycerate hydrolase, EC 4.2.1.11) were purified from both pig liver and muscle. Graphs of InC vs.r 2 from sedimentation equilibrium experiments are linear, which suggests homogeneous preparations of liver and muscle enolases. From these data the molecular weight of liver enolase is calculated to be approximately 92,000 D and that of muscle enolase to be approximately 85,000 D. SDS-PAGE experiments give a molecular weight value of 46,000 D for liver enolase and a value of 44,000 D for muscle enolase. These molecular weight values for liver and muscle enzymes are within the range for other enolases and show that both of these pig enolases are dimers. Amino acid composition data support the sedimentation equilibrium data and also give a smaller molecule weight (84,968 D) for muscle enolase compared to that of the liver enzyme (89,021 D). The two enzymes differ in their content of lysine [liver enolase (L)=94 residues, muscle enolase (M)=68 residues], histidine (L=13, M=21), serine (L=53, M=36), proline (L=52, M=34), and cysteine (L=4, M=21). Partial specific volumes of 0.737 ml/g for liver enolase and 0.735 ml/g for muscle enolase were calculated from the amino acid composition data. Pig liver and muscle enolases differ radically in their isoelectric points (pI=6.4–6.5 for liver enolase, and pI=8.8–9.0 for muscle enolase), and in their degree of inactivation by 750 mM LiCI (liver enolase is inactivated to a greater degree than the muscle enolase). Despite these physical and chemical differences, the kinetic constantsK M values for Mg2+, 2-phosphoglyceric acid, and phospho(enol)pyruvate appear not to be significantly different for these two forms of enolase. The physical, chemical, and kinetic data for pig liver and muscle enolases are compared to similar data for pig kidney enolase.  相似文献   

16.
Buccal juice of the sea hare Aplysia juliana was found to degrade algal polysaccharides. The optimal enzyme composition for protoplast preparation from Undaria pinnatifida was protein at 48 μg/ml buccal juice from sea hare, 10 mg/ml cellulase Onozuka-RS, 0.4 M NaCl, 0.8 M sorbitol, 2 mg/ml dextran sulfate sodium salt, and 1 μl/ml 2-mercaptoethanol in 10 mM MES buffer (pH 6.0). Protoplasts of Eisenia bicyclis, Endarachne binghamiae (Phaeophyta), and Ulva pertusa (Chlorophyta) could also be prepared in a similar manner. Yields of these protoplasts were about 107 cells per gram of fresh weight alga. Received January 26, 1998; accepted September 17, 1998.  相似文献   

17.
A trypsin inhibitor, termed ovostatin, has been purified approximately 265-fold with 82% yield, from unfertilized eggs of the sea urchin Strongylocentrotus intermedius, using trypsin coupled Sepharose 4B as an affinity column for chromatography. The isolated ovostatin is homogeneous in sodium dodecyl sulfate/polyacrylamide gel electrophoresis, the estimated molecular weight being 20K–21.5K. Ovostatin inhibits preferentially trypsin-like endogenous protease purified from the eggs of the same species and bovine pancreatic trypsin and also bovine pancreatic chymotrypsin. Values of IC50 (amount causing 50% inhibition of enzymes) for trypsin-like protease purified from eggs of the same species, bovine pancreatic trypsin, and bovine pancreatic chymotrypsin, are 0.91 ± 0.13 μg/ml (4.55 ± 0.65 × 10?8 M), 3.0 ± 0.28 μg/ml (1.5 ± 0.14 × 10?7 M), and 4.8 ± 0.2 μg/ml (2.4 ± 0.1 × 10?7 M), respectively, in the experimental condition used. Kinetic studies indicate that ovostatin is a noncompetitive inhibitor of trypsin. The inhibitor is relatively heat labile. NaCl (0.025–0.01 M) enhances the inhibitor activity, whereas KCl is inhibitory. Ovostatin requires a low concentration of Ca2+ for activity. The activity is higher in unfertilized eggs than in fertilized eggs; total activity and specific activity in unfertilized eggs is about 1.67-fold and 1.85-fold higher than those in fertilized eggs, respectively. We believe that ovostatin may regulate the function of the cortical granule protease and other trypsin-like proteases that are activated in sea urchin eggs during fertilization.  相似文献   

18.
The renaturation kinetics of mitochondrial DNA from the yeast Saccharomyces carlsbergensis have been studied at different temperatures and molecular weights. At renaturation temperatures 25 deg. C below the mean denaturation temperature (Tm) in 1 M-sodium chloride the renaturation rate constant is found to decrease with increasing molecular weight of the reacting strands. This unusual molecular weight dependency gradually disappears with an increase in the renaturation temperature. At a temperature 10 deg. C below the melting point, the rate constant shows the normally expected increase with the square root of the molecular weight. From the renaturation data at this temperature, the molecular weight of the mitochondrial genome is estimated to be about 5·0 × 107. The same size of genome was found from renaturation at low molecular weight and 25 deg. C below the Tm.The sedimentation properties of denatured mitochondrial DNA at pH values 7·0 to 12·5 were used to study the conformation of this DNA in 1 M-sodium chloride. The results obtained support the conclusion from the renaturation studies: that the pieces of denatured mitochondrial DNA with a molecular weight above 2 × 105 to 3 × 105, in 1 M-sodium chloride at 25 deg. C below the mean denaturation temperature are not fully extended random coils. Presumably, interaction between adenine and thymine-rich sequences, which are clustered at certain distances within the molecules, is the molecular basis for these observations.  相似文献   

19.
Cathepsin B (EC 3.4.22.1) from rat liver was crystallized and its amino acid composition was determined. The purified enzyme formed spindle-shaped crystals and its homogeneity was proved by ultracentrifugical analysis. Its S20, W value was 2.5 S and its molecular weight was calculated to be 24,000 from the result of sedimentation equilibrium analysis. Amino acid analysis showed that it contained glucosamine and galactosamine. The activity of the protease was maximal at pH 6.0 with α-N-benzoyl-DL-arginine p-nitroanilide as substrate. The apparent Kms for α-N-benzoyl-DL-arginine p-nitroanilide and α-N-benzoyl-DL-arginine-2-naphthylamide were 1.4 × 10?2 M and 2.0 × 10?3 M, respectively  相似文献   

20.
The viscosity of bovine liver glutamate dehydrogenase solutions was studied at 10 and 20° C in 0.2.M sodium phosphate buffer at pH 7, in the concentration range 0.1–8 mg/ml. A method for the study of the viscosity of very dilute solutions of associating enzymes is described. It was found that the reduced specific viscosity ηsp/c of glutamate dehydrogenase continuously increases with increasing enzyme concentration, from about 4 ml/g at the lowest concentrations to about 16 ml/g at 8 mg/ml. In the presence of 10?3M GTP and 10?3M NADH the viscosity increase is much smaller and the results can be extrapolated to zero enzyme concentration to yield an intrinsic viscosity [η] = 3.2. The values of ηsp/c in phosphate buffer alone apparently extrapolate to the same value of [η], or to a value close to it. We also observe that, in the presence of toluene ηsp/c increases very much more with enzyme concentration: ηsp/c already equals 16 ml/g at a concentration of 0.75 mg/ml. These observations are in good agreement with the hypothesis that the active oligomer of glutamate dehydrogenase (MW = 312,000) associates with increasing enzyme concentration to form linear rodlike polymers of indefinite length. This association is strongly diminished by the addition of 10?3M GTP, 10?3M NADH. Toluene, on the other hand, promotes reversible association to linear polymers of very high molecular weight. The transverse and axial rotary frictional coefficients of macroscopic bodies, similar to a physical model for the structure of glutamate dehydrogenase recently advanced, were determined. Assuming that the viscosity of the model is equal to that of an ellipsoid of rotation with identical frictional coefficients, we calculate [η] = 3.26 ml/g according to Kuhn and 3.20 ml/g according to Simha, for the glutamate dehydrogenase oligomer, in good agreement with the result derived from the study of enzyme solutions.  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