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.
Lina Zhang  Liqun Yang 《Biopolymers》1995,36(6):695-700
A Water-soluble glucan A was isolated from the fruit body of Auricularia auricula-judae. It is composed of a backbone chain of β-(1 → 3)-linked D -glucose residues, two out of three glucose residues being substituted at C-6 positions with a single glucose unit. The weight average molecular weight Mw, number average molecular weight Mn, and intrinsic viscosity [η] of the fractionated samples were studied at 25°C in water and in dimethylsulfoxide (DMSO). The Mark-Houwink equation was established as [η] = 6. 10 × 10?4 Mw1.14 for the glucan A having Mw ranging from 9 × 105 to 1.6 × 106 in water. The values of [η] in water are far higher than those in DMSO, but the values of Mn measured in water are the same as those in DMSO. Analysis of Mw and [η] in terms of the known theories for rods and wormlike chains yielded 1030 ± 100 nm?1, 90 ± 20 nm, 1.3 ± 0.3 nm, and 0.26 ± 0.03 nm for molar mass per unit contour length ML, persistence length q, diameter d, and contour length h per main-chain glucose residue, respectively. The present data suggest that glucan A dissolves in water as single-stranded helical chains and in DMSO as semiflexible chains. © 1995 John Wiley & Sons, Inc.  相似文献   

2.
K Gekko  H Noguchi 《Biopolymers》1971,10(9):1513-1524
The number average molecular weight, Mn, of low molecular weight dextran was determined through endgroup analysis, and the intrinsic, viscosities of these materials in aqueous solution were determined at 25°C. The ultrasonic velocities in their aqueous solutions were also measured at 25 and 45°C. As concerns the molecular weight dependence of the intrinsic viscosity, partial specific compressibility of solute and the hound water around the solute, the following results were obtained. (1) log [η]-log Mn and [η]/Mn0.5 – Mn0.5 plots were in accord with the Mark-Houwink and Stockmayer-Fix-man equations respectively for Mn > 2, 000, but these plots deviated from the equations for Mn < 2, 000. (2)The partial specific compressibility, β 1°, of dextran is expressed by following equation for Mn < 2,000: β 1° = 10?12 × (13.6 log Mn - 51.7) (cm2/dyne). In contrast, it, becomes the constant value, -- 7.3 × 10?12 cm2/dyne, for Mn > 2,000. (3) The amount of bound water of dextran calculated from the sound velocity measurement lakes constant value of 0.17 ml g for Mn > 2, 000, but the amount of hydration increase with decreasing molecular weight for Mn < 2,000. From these results, a dextran molecule in aqueous solution is expected to change its conformation from random coiling to uncoiling stretched form at the molecular weight of around 2, 000 or about 12 glucose units.  相似文献   

3.
Molecular characteristics for pullulan, a polysaccharide produced by a fungus Aureobasidium pullulans, were measured by light scattering, viscometry, and gel-permeation chromatography. From the experimental data the Mark-Houwink-Sakurada viscosity equation in water at 25°C was determined for samples having the molecular weight M ranging from 48 × 103 to 2.18 × 106 g mol?1 as [η] = (1.91 ± 0.02) × 10?2Mw0.67±0.01 (in cm3 g?1); and as molecular weight decreased, the slope of the viscosity equation decreased, although the molecular weight values below 30 × 103 g mol?1 evaluated by gel-permeation chromatography were somewhat unreliable. The unperturbed dimensions 〈R201/2 of pullulan were estimated by determining the expansion factor αs, from the theoretical combination of theories for the interpenetration function Ψ and those for αs. The 〈R20/M value estimated from this procedure in 6.7 × 10?17 cm2 mol g?1. We concluded that the polysaccharide chain that is linked by the α-1,6-glucosidic linkage behaves like a flexible chain in aqueous solution.  相似文献   

4.
The intrinsic viscosity and sedimentation coefficient, of native and single-stranded T7 DNA have been determined at 25°C as a function of ionic strength in neutral and alkaline NaCl. The relationship between [η] and S,w is well represented by the Mandelkern-Flory equation over the entire range of conditions between 0.0013 and 1M Na+. An apparent discrepancy between the two methods at moderate to high ionic strengths is probably due to a change in V with ionic strength. It appears that [η] is a more sensitive and reliable measure of molecular expansion for native DNA, S,w but is a better index of conformational change in single strands, since [η] becomes too small to measure conveniently at high ionic strengths. At moderate to high ionic strengths, denaturation leads to a decrease in [η], although unfolded single strands retain considerable viscosity. At sufficiently low ionic strength, the intrinsic viscosity of the single strands becomes higher than that of native DNA, and the effective volume of a single strand approaches that of the native molecule.  相似文献   

5.
Molecular weight and shape of the phycocyanin hexamer   总被引:1,自引:0,他引:1  
M Kato  W I Lee  B E Eichinger  J M Schurr 《Biopolymers》1974,13(11):2293-2304
The hexamer of phycocyanin from Phormidium luridum has been isolated and purified by ammonium sulfate fractionation and gel chromatography. The protein is characterized by the sedimentation constant S°20, w = 10.2S, the diffusion coefficient D20, w = 4.73 × 10?7 cm2/sec, and intrinsic viscosity [η] = 3.89 ml/g. The molecular weight of the aggregate is 209,000. The shape and dimensions of the hexamer are discussed in terms of a model consisting of subunits arranged with C6 symmetry. The monomers, assumed to be spherical, are found to have a radius of 22 Å, and the diameter across the hexamer is 132 Å. The latter figure agrees closely with dimensions observed in electron micrographs.  相似文献   

6.
Simple approximate expressions have been derived from the theory of Zimm and Bragg for use in the analysis of experimental data on the helix-coil transition in polypeptide. On the basis of the resulting expressions practical procedures are proposed to determine two basic parameters characterizing a thermally induced transition, i.e., helix initiation parameter σ and enthalpy change for helix formation, ΔH. They have been applied to the data for poly(β-benzyl L -aspartate) (PBLA) with the result: σ = 1.6 × 10?4 and ΔH = ?450 cal/mole for PBLA in m-cresol; σ = 0.6 × 10?4 and ΔH = 260 cal/mole for PBLA in chloroform containing 5.7 vol-% of dichloroacetic acid. This result gives evidence that σ may change not only from one polypeptide to another but also for a given polypeptide in different solvents. The change in limiting viscosity number [η] accompanying the transition was measured in the same solvents. The curve of [η] versus helical content had a relatively monotonic shape for the chloroformdichloroacetic acid solutions as compared with that for the m-cresol solutions, indicating that [η] depended largely on σ. Provided that [η] is a direct measure of the mean-square radius of gyration, 〈S2〉, the results are consistent with the theoretical predictions of Nagai and of Miller and Flory for 〈S2〉.  相似文献   

7.
The z‐average mean‐square radius of gyration 〈S2z, the particle scattering function P(k), the second virial coefficient, and the intrinsic viscosity [η] have been determined for amylose tris(phenylcarbamate) (ATPC) in methyl acetate (MEA) at 25°C, in ethyl acetate (EA) at 33°C, and in 4‐methyl‐2‐pentanone (MIBK) at 25°C by light and small‐angle X‐ray scattering and viscometry as functions of the weight‐average molecular weight in a range from 2 × 104 to 3 × 106. The first two solvents attain the theta state, whereas the last one is a good solvent for the amylose derivative. Analysis of the 〈S2z, P(k), and [η] data based on the wormlike chain yields h (the contour length or helix pitch per repeating unit) = 0.37 ± 0.02 and λ?1 (the Kuhn segment length) = 15 ± 2 nm in MEA, h = 0.39 ± 0.02 and λ?1 = 17 ± 2 nm in EA, and h = 0.42 ± 0.02 nm and λ?1 = 24 ± 2 nm in MIBK. These h values, comparable with the helix pitches (0.37–0.40 nm) per residue of amylose triesters in the crystalline state, are somewhat larger than the previously determined h of 0.33 ± 0.02 nm for ATPC in 1,4‐dioxane and 2‐ethoxyethanol, in which intramolecular hydrogen bonds are formed between the C?O and NH groups of the neighbor repeating units. The slightly extended helices of ATPC in the ketone and ester solvents are most likely due to the replacement of those hydrogen bonds by intermolecular hydrogen bonds between the NH groups of the polymer and the carbonyl groups of the solvent. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 729–736, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

8.
A simple purification method which enables us to obtain homogeneous proteinase C from S. cerevisiae was developed. Physical and chemical properties of the purified enzyme were determined. The extinction coefficient at 280 mμ, , of yeast proteinase C was 14.8, and its isoelectric point was pH 3.60. Partial specific volume, intrinsic viscosity and the sedimentation and diffusion coefficients of homogeneous protein were , 0.71 ml/g, [η], 4.83 × 10?2ml/g, , 4.23 S and , w, 6.1 × 10?7 cm2/sec. From these values, molecular weights, M[·],D, MS,D and M[·],S, of 60,000, 59,000 and 58,000, respectively, were obtained. The sedimentation equilibrium experiment gave a molecular weight, Mequil, of 61,000. Yeast proteinase C contained 11.9% nitrogen and was a glycoprotein with 16.7% carbohydrate: The value of β-function, 2.163×l06 or 2.20×l06 indicates that the molecular shape of yeast proteinase C is a plorate with an axial ratio of 4.0, assuming 35% hydration. Furthermore, yeast proteinase C may be a compact, asymmetric ellipsoidal model having semi-axes 30Å × 30Å × 130Å.  相似文献   

9.
A new oscillating crucible viscometer, having a U-shaped circular channel, is described. The damping coefficient δ is lowered by an increase of the viscosity η. The instrument described here allows the solution to come in contact with inert plastic only. At all steps of its preparation and during viscosity measurements, giant DNA from rat liver nuclei was maintained at shear stresses around 10?4 dynes cm?2. Viscosity was studied as a function of surface tension, DNA concentration and shear stress. It was found that under our experimental conditions it was possible to obtain meaningful values for reduced viscosity, ηred, practically identical to intrinsic viscosity [η]. Rat liver nuclei are incubated in an alkaline lysing solution (pH 12.5; 22 °C): they are lysed immediately and the released DNA starts to uncoil. The viscosity of solutions of this giant DNA increases very slowly with time, reaching a maximum only after about ten hours. The process was accelerated by single-stranded breaks arising from methylation of DNA in vivo with dimethylnitrosamine. It was found that the time of DNA disentanglement was sensitive to an exceedingly small number of breaks. We think that we were able to measure molecular weights around the length of the single strand of an average chromosome (Mn 5 × 1010). An empirical relation between molecular weight and reduced viscosity after complete disentanglement was also established, as a linear log-log plot, covering a molecular weight range between 108 and 2.5 × 1010. It is suggested that the viscosimetric evaluation of DNA disentanglement is probably the most sensitive method for studying DNA damage induced “in vivo” by chemical carcinogens.  相似文献   

10.

In the present work, a positive effect was obtained by using low molecular weight chitosan compounds in combination with synthetic fungicides. Antifungal activity against Botrytis cinerea, determined by the radial growth method, was more than 75%, with a 25?×?10??10 g/L concentration of fludioxonil or difenoconazole in compounds. Metabolic activity of B. cinerea fungus was about 15% when using a chitosan compound containing fludioxonil at a concentration of 25?×?10??7 g/L. The combined action of chitosan with difenoconazole at a fungicide concentration of 25?×?10??4 g/L is 2–3 times more effective than the action of each component separately. Results of studies for artificially inoculated B. cinerea tomato fruit when treated with low molecular chitosan and chitosan conjugate with gallic acid reduced the frequency of rotting fruit by 50 and 83%, respectively. Chitosan-gallic acid conjugate were obtained from chitosans with Mw of 28 kDa (Ch28GA) was proved to be effective as a preventive treatment for 3 days and can potentially be used as a biofungicide against B. cinerea on tomatoes in the post-harvest period.

  相似文献   

11.
Purified, water-soluble polysaccharide of Phellodendron amurense Ruprecht (P-WSPS) was obtained in highly homogeneous state. The partial specific volume and intrinsic viscosity were, respectively, 0.56 and 11.68 dL/g. The S value and molecular weight depended strongly on the P-WSPS concentration, but extrapolation to infinite dilution gave Sw,200 = 10.67 × 10?13 s?1 and M0 = 625 × 103 from velocity-sedimentation and equilibrium-sedimentation experiments, respectively, results that showed the highly branched structure of P-WSPS. The molecular weight calculated from the D0 value (1.035 × 10?7) and S0 value agreed with that from the equilibrium-sedimentation experiment to within ~±9%, showing that the diffusion method is useful for rapid estimation of molecular weight.  相似文献   

12.
Given the increase in the incidence of insulin resistance, obesity, and type 2 diabetes in children and adolescents, it would be of paramount importance to assess quantitative indices of insulin secretion and action during a physiological perturbation, such as a meal or an oral glucose‐tolerance test (OGTT). A minimal model method is proposed to measure quantitative indices of insulin secretion and action in adolescents from an oral test. A 7 h, 21‐sample OGTT was performed in 11 adolescents. The C‐peptide minimal model was identified on C‐peptide and glucose data to quantify indices of β‐cell function: static φs and dynamic φd responsivity to glucose from which total responsivity φ was also measured. The glucose minimal model was identified on glucose and insulin data to estimate insulin sensitivity, SI, which was compared to a reference measure, SIref, provided by a tracer method. Disposition indices, which adjust insulin secretion for insulin action, were then calculated. Indices of β‐cell function were φs = 51.35 ± 8.89 × 10?9min?1, φd = 1,392 ± 258 × 10?9, and φ = 82.09 ± 17.70 × 10?9min?1. Insulin sensitivity was SI = 14.19 ± 2.73 × 10?4, not significantly different from SIref = 14.96 ± 3.04 × 10?4 dl/kg·min per µU/ml, and well correlated: r = 0.98, P < 0.0001, thus indicating that SI can be accurately measured from an oral test. Disposition indices were DIs = 1,040 ± 201 × 10?14 dl/kg/min2 per pmol/l, DId = 33,178 ± 10,720 × 10?14 dl/kg/min per pmol/l, DI = 1,844 ± 522 × 10?14 dl/kg/min2 per pmol/l. Virtually the same minimal model assessment was obtained with a reduced 3 h, 9‐sample protocol. OGTT interpreted with C‐peptide and glucose minimal model has the potential to provide novel insight regarding the regulation of glucose metabolism in adolescents, and to evaluate the effect of obesity and interventions such as diet and exercise.  相似文献   

13.

A new exopolysaccharide (EPS) producing Gram-positive bacterium was isolated from the rhizosphere of Bouteloua dactyloides (buffalo grass) and its EPS product was structurally characterized. The isolate, designated as LB1-1A, was identified as Bacillus paralicheniformis based on 16S rRNA gene sequence and phylogenetic tree analysis. The EPS produced by LB1-1A was identified as a levan, having β(2?→?6) linked backbone with β(2?→?1) linkages at the branch points (4.66%). The isolate LB1-1A yielded large amount (~?42 g/l) of levan having high weight average molecular weight (Mw) of 5.517?×?107 Da. The relatively low degree of branching and high molecular weight of this levan makes B. paralicheniformis LB1-1A a promising candidate for industrial applications.

  相似文献   

14.
Quartz crystal microbalance (QCM) biosensor was used for the chiral recognition of five pairs of enantiomers by using goat serum albumin (GSA) and rabbit serum albumin (RbSA) as chiral selectors. Serum albumin (SA) was immobilized on the QCM through the self‐assembled monolayer technique, and the surface concentration of GSA and RbSA were 8.8 × 10?12 mol cm?2 and 1.2 × 10?11 mol cm?2, respectively. The QCM biosensors showed excellent sensitivity and selectivity. Meanwhile, the chiral recognition of SA sensors was quite species dependent. There were differences between GSA and RbSA sensors in the ability and the preference of chiral recognition. To R,S‐1,2,3,4‐tetrahydro‐1‐naphthylamine (R,S‐1‐TNA), R,S‐1‐(4‐methoxyphenyl)ethylamine (R,S‐4‐MPEA), and R,S‐1‐(3‐methoxyphenyl)ethylamine (R,S‐3‐MPEA), the preference of the stereoselective SA‐drug binding of the two kinds of SA sensors were consistent. However, to R,S‐2‐octanol (R, S‐2‐OT) and R,S‐methyl lactate (R,S‐MEL), the two kinds of SA sensors had opposite chiral recognition preference. Moreover, the interactions of SA and the five pairs of enantiomers have been further investigated through ultraviolet (UV) and fluorescent (FL) spectra. The UV/FL results were in accordance with the consequence of QCM. Chirality 24:804–809, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

15.
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   

16.
17.
The real and imaginary parts of complex viscosity, η′ and η″, are measured for dilute solutions of poly(γ-benzyl-L -glutamate) in m-cresol, a helicogenic solvent. The frequency range is 2.2–525 kHz; the concentration range 0.2–5 g/dl; the temperature 30°C, and the molecular weights Mr are 6.4 × 104–17 × 104. The dispersion curve of extrapolated intrinsic dynamic viscosity [η′] of samples with Mr > 105 is interpreted in terms of three mechanisms appearing from low to high frequencies: end-over-end rotation, flexural deformation, and side-chain motion. For a sample with Mr < 105, the flexural relaxation disappears and a plateau of [η′] is distinctly observed between rotational and side-chain relaxations. Rotational relaxation times of all the samples obey the Kirkwood–Auer theory. The strong concentration dependence of rotational relaxation time is explained by collisions of molecules rather than association. Flexural relaxation times calculated from a theory by assuming the persistence length as 1200 Å are consistent with observed dispersion curves of [η′].  相似文献   

18.
19.
The coupling reaction of glucoamylase and halogenacetyl cellulose (HAC) without pretreating with organic solvent led to form large particles of immobilized glucoamylase and the activity and the specific activity of the preparation were very low. However, the coupling reaction with HAC pretreated with organic solvents allowed to form very fine particles and the activity was increased by five times. The latter contained 3~6% of enzyme protein and the specific activity to maltose reached to 80~90% of native glucoamylase. Since the specific activity of the preparation was presumed to be much influenced by the particle size, the specific activity and general properties of different particle sizes were compared with those of native enzyme. The specific activities of particles of 0~15μ, 15~55μ, 70~190μ and 130~270μ showed 82%, 33%, 27% and only 7% of native enzyme, respectively. Km values of native form, 0~15μ, 15~55μ and 70~190μ particles were 0.90×10?3m, 1.35×10?3m, 1.60×10?3m and 2.1×10?3m in the case of maltose as substrate, respectively. The other properties of particles of 0~15μ were almost identical to those of native enzyme except for the effect of temperature on the reaction rate. However, pH activity, pH stability and urea stability of particles of 70~190μ were much inferior to those of native enzyme and particles of 0~15μ.  相似文献   

20.
Interactions of Isatin and its derivatives, Isatin-3-isonicotinylhydrazone (IINH) and Isatin-β-thiosemicarbazone (IBT), with calf thymus DNA (ctDNA) have been investigated to delineate pharmaceutical-physicochemical properties using UV–Vis/fluorescence/circular dichroism (CD) spectroscopy, viscosity measurements, and multivariate chemometrics. IINH and IBT molecules intercalate between base pairs of DNA, hypochromism in UV absorptions, increase in the CD positive band, sharp increase in specific viscosity, and the displacement of the methylene blue and Neutral Red dye in complexes with ctDNA, by the IINH and IBT molecules, respectively. The observed intrinsic binding constants (Kb[IBT–ctDNA]?=?1.03 × 105 and Kb[IINH–ctDNA]?=?1.09 × 105 L mol?1) were roughly comparable to other intercalators. In contrast, Isatin binds with ctDNA via groove mode (Kb[Isatin–ctDNA]?=?7.32 × 104 L mol?1) without any significant enhancement in ctDNA viscosity. The fluorescence quenching of Isatin by ctDNA was observed as static. CD spectra indicated that Isatin effectively absorbs into grooves of ctDNA, leading to transition from B to C form. Thermodynamic parameters like enthalpy changes (?H < 0) and entropy changes (?S > 0) were calculated according to Van’t Hoff’s equation, indicating the spontaneous interactions. The common soft/hard chemometric methods were used not only to resolve pure concentration and spectral profiles of components using the acquired spectra but also to calculate Stern–Volmer quenching constants, binding stoichiometry, apparent binding constants (Ka), binding constants (Kb), and thermodynamic parameters. The Kb values for Isatin, IINH, and IBT were calculated as 9.18 × 103, 1.53 × 105, and 2.45 × 104 L mol?1, respectively. The results obtained from experimental-spectroscopic analyses showed acceptable agreement with chemometric outlines.  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