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.
The phase diagram of the guar–dextran aqueous two-phase system has been described on the basis of static light scattering measurements in the dilute regime. By determining the molar weight and second virial coefficient from the two single polymers and the second virial cross coefficient from mixtures at constant guar/dextran ratio (either 17/83 or 28/72), the thermodynamic models based on the virial expansion or the Flory–Huggins theory were successfully applied. The second virial coefficient of guar was difficult to estimate with enough accuracy by light scattering and therefore was obtained by adjustment using a simple criterion stating that the calculated spinodal passes through the experimental critical point. The obtained value was within the confidence interval given by light scattering. Virial expansion and Flory–Huggins approaches yielded quite similar theoretical phase diagrams that satisfactorily fitted the experimental one. The slight discrepancies observed on the position of binodals and critical points has been attributed to the polydispersity of guar or the difficulty in extrapolating from the dilute regime to the semidilute one. The slope of the tie-lines was predicted with a good accuracy, especially with the virial expansion model. The fact that both approaches gave such similar results is probably related to the fact that the expressions of chemical potentials are equivalent if the polymer concentrations are low enough. In this particular case, both models are based on excluded volume interactions and equally describe the phase behavior of the guar–dextran aqueous system.  相似文献   

2.
The sol–gel transition in aqueous alginate solutions of four alginate samples having different molecular weights (MW) and M/G ratios induced by cupric cations was monitored by rheology measurements. The gel point fgel and the relaxation critical exponent n were determined using the Winter’s criterion over the alginate concentration CAlg of 1–4 wt%. The scaling for the zero shear viscosity η0 before the gel point and the equilibrium modulus Ge after the gel point was established against the relative distance ε from the gel point at the concentration of CAlg = 1 wt%, giving the critical exponents k and z. The results indicated that fgel was almost independent of the alginate concentration and became higher for the sample with lower molecular weight. The critical exponent n decreased with the increase in CAlg for these four Cu-alginate samples and the fractal dimension df estimated from n suggested a denser structure in the critical gel with high G content. The critical exponent n evaluated from k and z agreed well with n determined from the Winter’s criterion.  相似文献   

3.
A rapid method for establishing the constants in the Mark–Houwink–Sakurada equation, relating intrinsic viscosity and molecular weight (MW), of guar galactomannan is described. Following partial acid hydrolysis, the galactomannan was analyzed using high performance size exclusion chromatography employing viscosimetry and right angle light scattering detectors. In this way, a large number of samples of polysaccharides with a wide range of MW distributions were prepared, without need for isolation, and intrinsic viscosity and MW rapidly determined. The a and K values found for guar galactomannan were 0.72 and 5.13×10−4 ([η] in dl/g) respectively, in good agreement with previously published values.  相似文献   

4.
A method for determination of carboxymethyl cellulose (CMC) molecular weight (MW) and chemical heterogeneity (degree of oxidation (DO)) using a bi-detector HPSEC (UV-detector online with refractometer) has been developed. It has been found that the use of 0.5 N NaOH or 0.4 M acetate buffer as the eluent ensures CMC separation according to MW. It has been revealed that the universal calibration for the polyelectrolyte CMC and the neutral polymer dextran is valid under the conditions applied. The Mark–Houwink equations for CMC in 0.5 N NaOH and 0.4 M acetate buffer have been estimated to be [η]=5.37×10−4 MW0.73 and [η] =2.24×10−4 MW0.83 dl g−1, respectively. The equation log K=1.64−4.00 ml g−1 for CMC has been estimated. An approach for determining DO from adsorption at 290 or 313 nm has been developed.  相似文献   

5.
Vortex flow reactors (VFRs) are a good option when fragile particles are present in the medium, due to their gentle but efficient stirring characteristics. However, the presence of a by-pass stream may deteriorate the reactor performance, and particles of inadequate density may either settle down or clog the reactor outlet. This work assessed the performance of an enzymic VFR. Fructose–glucose isomerization, catalyzed by immobilized glucoisomerase was the test reaction, taking advantage of the negligible changes that it causes on the medium viscosity. Intra- and extra-particle mass transfer effects were avoided. Reactor geometry (radius ratio η=0.677 and aspect ratio Γ=18.30) and residence time were selected aiming at possible applications of the device as a bioreaction and/or adsorption system. Visualization experiments confirmed that the vortices’ cores stop their axial displacement when the rotation of the inner cylinder is increased. Intermediate rotations were the most detrimental to reactor performance, due to by-pass effects. Vortex agitation is very gentle, causing no detectable damage to shear-sensitive particles.  相似文献   

6.
Both dilute and concentrated solutions of bacterial Mannuronan (MANNA) and its epimerized products by AlgE1 at 5 and 24 h, named MANNAEp1t5h and MANNAEp1t24h, respectively, and AlgE4 (MANNAEp4) have been studied as a function of variables such as polymer concentration and ionic strength (NaCl) in order to investigate the macromolecular solution properties of these innovative polyuronic acids having the same charge density but different composition and sequence of β-d-mannuronic acid (-M-), -l-guluronic acid (-G-) or MG-blocks.

Measurements of intrinsic viscosity [η] as a function of ionic strength, I, by capillary viscometry has led to an estimate of the Smidsrød-Haug parameter B, an index useful to characterize the stiffness of polymeric chains. The results are largely consistent with much of the published data relative to chain extension and conformational freedom around the torsional angles of the glycosidic linkages occurring in alginates.

Steady shear rheometry provided information about the coil-overlapping parameter c*, which marks the transition from dilute to concentrated solution. The slopes of the double logarithmic plots of ηsp vs. c[η] both at low and high degrees of coil overlap suggest that all samples solutions behave like linear polymer entangled network systems. The value of c* is strictly influenced by the stiffness of the chains, and hence by the primary structure.

Dynamic shear rheometry shows that the frequency dependence of dynamic viscosity is only partially superimposable to the shear rate dependence of viscosity. Such behaviour may be ascribed to the presence of semiflexible polymeric coils in a non-totally destructured entangled state.

By solvent/non-solvent (H2O/isopropanol) fractionation carried out on mannuronan, a set of samples with different average molecular weights and narrow polydispersities were obtained. Triple detection GPC allowed the evalutation of the Mark–Houwink–Sakurada parameters as well as of the characteristic ratio C for one of the fractionated MANNA samples. The chain persistence length was estimated by the wormlike chain model.  相似文献   


7.
Human estrogenic 17β-hydroxysteroid dehydrogenase (17β-HSD1, EC1.1.1.62) is an important enzyme that catalyses the last step of active estrogen formation. 17β-HSD1 plays a key role in the proliferation of breast cancer cells. The three-dimensional structures of this enzyme and of the enzyme-estradiol complex have been solved (Zhu et al., 1993, J. Mol. Biol. 234:242; Ghosh et al., 1995, Structure 3:503; Azzi et al., 1996, Nature Struct. Biol. 3:665). The determination of the non-reactive ternary complex structure, which could mimic the transition state, constitutes a further critical step toward the rational design of inhibitors for this enzyme (Ghosh et al. 1995, Structure 3:503; Penning, 1996, Endocrine-Related Cancer, 3:41).

To further study the transition state, two non-reactive ternary complexes, 17β-HSD1–EM519-NADP+ and 17β-HSD1–EM553-NADP+ were crystallized using combined methods of soaking and co-crystallization. Although they belong to the same C2 space group, they have different unit cells, with a=155.59 Å, b=42.82 Å, c=121.15 Å, β=128.5° for 17β-HSD1–EM519-NADP+, and a=124.01 Å, b=45.16 Å, c=61.40 Å, β=99.2° for 17β-HSD1–EM553-NADP+, respectively. Our preliminary results revealed that the inhibitors interact differently with the enzyme than do the natural substrates.  相似文献   


8.
Static light scattering of high amylopectin waxy maize starch gently dispersed in 90% dimethyl sulfoxide–water yielded a weight average molecular weight Mw and radius of gyration Rg of 560×106 g/mol and 342 nm, respectively. To obtain an independent hydrodynamic characterization of these solutions, we measured the sedimentation coefficient for the main component in an analytical ultracentrifuge. The value of s0, the infinite dilution sedimentation coefficient, was 199 S. The translational diffusion coefficient D0 in very dilute solutions was measured by dynamic light scattering at 90° and found to be 2.33×10−9 cm2/s. An effective hydrodynamic radius Rh was calculated from this diffusion constant using the Stokes–Einstein equation and found to be 348 nm. The structure-related parameter ρ=Rg/Rh was calculated to be 0.98. The weight average molecular weight calculated from the Svedberg equation using the values measured for s0 and D0 was 593×106 g/mol. This result is in reasonable agreement with the light scattering results. As light scattering results are subject to experimental errors due to the possibility of dust contamination, the presence of microgel or aggregates, and the questionable applicability of light scattering theory to interpret results for macromolecular sizes approaching the wave length of light used as a source for scattering, it is advisable to have corroborating hydrodynamic data when possible to further validate light scattering results in this very high molecular weight range.  相似文献   

9.
Guar gum/poly(acrylic acid) semi-interpenetrating polymer network (IPN) hydrogels have been prepared via free radical polymerization in the presence of a crosslinker of N,N′-methylene bisacrylamide (MBA). The kinetics of swelling and the water transport mechanism were studied as a function of the composition of the hydrogels and the pH of the swelling medium. Hydrogels showed enormous swelling in aqueous medium and displayed swelling characteristics, which were highly dependent on the chemical composition of the hydrogels and pH of the medium in which hydrogels were immersed (ionic strength I = 0.15 mol/L). The semi-INP hydrogels were characterized by evaluating various network parameters such as average molecular weight between crosslinks (Mc) crosslink density (ρ) and mesh size ξ.  相似文献   

10.
Synthesis of chitosan sulfates with low molecular weight (Mv 9000–35,000 Da) was carried out by sulfation of low molecular weight chitosan (Mv 10,000–50,000 Da). The oleum was used as sulfating agent and dimethylfornamide as medium. The chitosans were prepared by enzymatic and acidic hydrolysis of initial high molecular weight chitosan as well as by extrusion solid-state deacetylation of chitin. As was shown by FT-IR and NMR-methods and elemental analysis, the sulfation occurred at C-6 and C-3 positions and substitution degree is 1.10–1.63. The molecular weight sulfated chitosan was determined by viscometric method and the Mark–Houwink equation [η]=10−5 4.97 M0.77. Study of anticoagulant activity showed that chitosan sulfates with lowered molecular weight demonstrated a regular increase of anti-Xa activity like heparins.  相似文献   

11.
Shear flow, dynamic oscillation and extensional viscosity measurements were used to compare the rheological performance of several hylan samples (Mv 1.6, 3.2, 3.7, 4.7 and 5.6×106) and hyaluronan (Mv 1.4 and 1.8×106) before and after hydroxyl radicals (√OH) induced degradation. It was found that the higher molecular weight cross-linked structure of hylan was more resistant to degradation than hyaluronan and that this superior stability was reflected in various rheological parameters. The √OH degradation of the initial hylan and hyaluronan samples produced a range of polysaccharides based on hylan and hyaluronan with molecular weight covering a range from 0.5–5.6×106. The rheological parameters associated with the polysaccharides could then also be studied. Zero shear values of the complex viscosity (η*), dynamic viscosity (η′) and shear viscosity (η) were calculated using the method of Morris1 and shown to approach the same value at zero shear or frequency. An adaptation of the method of Gibbs et al.2 gave a ‘master curve’ for the storage and loss modulus of hyaluronan and hylan, which encompasses a 10-fold molecular weight and a 5-fold concentration variation. In all instances for hylan, the storage modulus predominates over the loss modulus, whereas for hyaluronan, the reverse is true, demonstrating the greater elasticity of hylan throughout the whole experimental range of molecular weights and concentrations.  相似文献   

12.
The hydrodynamic properties (intrinsic viscosity, [η]; infinite dilution sedimentation coefficient, s20,w0; weight average molecular weight, Mw and translational frictional ratio, f/f0) of a high methoxy pectin have been evaluated at various temperatures (20–60°C). A reduction in the value of all four hydrodynamic parameters is indicative of depolymerisation and is in agreement with an earlier study using viscometry [Axelos, M.A.V., & Branger, M., (1993). Food Hydrocolloids, 7, 91–102]. The apparent linearity of the Mark – Houwink plot of log[η] vs log Mw suggests that the conformation of the pectin molecule does not change significantly over the temperature range studied. The evaluation of the Mark–Houwink viscosity exponent (a=0.84) indicates a moderately extended structure. This then allows the calculation of the number of Kuhn statistical lengths per chain from the adapted ‘blob’ theory of Dondos [Dondos A. (2001). Polymer, 42, 897–901]. The weight average number of Kuhn statistical lengths per chain is reduced from (170±10) to (125±10) when the temperature is increased from 20–60°C. This may be of significance as many high methoxy pectins are exposed to high temperatures during processing in both the food and pharmaceutical industries.  相似文献   

13.
This study was conducted to determine the utility of deletion spectrum and mutant frequency (MF) of the hypoxanthine phosphoribosyl transferase gene (HPRT) as indicators of radiation exposure in Russian Liquidators who served in 1986 or 1987 in the clean up effort following the nuclear power plant accident at Chernobyl. HPRT MF was determined using the cloning assay for 117 Russian Controls and 122 Liquidators whose blood samples were obtained between 1991 and 1998. Only subjects from whom mutants were obtained for deletion analysis are included. Multiplex PCR analysis was performed on cell extracts of 1080 thioguanine resistant clones from Controls and 944 clones from Liquidators. Although the deletion spectra of Liquidators and Controls were similar overall, the Liquidator deletion spectrum was heterogeneous over time. Most notable, the proportion of total gene deletions was higher in 1991–1992 Liquidators than in Russian Controls (χ2=10.5, p=0.001) and in 1993–1994 Liquidators (χ2=8.3, p=0.004), and was marginally elevated relative to 1995–1996 Liquidators (χ2=3.3, p=0.07). This type of mutation has been highly associated with radiation exposure. Total gene deletions were not increased after 1992. Band shift mutations were also increased in the 1991–1992 Liquidators but were associated with increased MF of both Liquidators and Controls (p=0.009), not with increased MF in 1991–1992 Liquidators (p=0.7), and hence are not believed to be associated with radiation exposure. Regression analysis demonstrated that relative to Russian Controls HPRT MF was elevated in Liquidators overall when adjusted for age and smoking status (37%, p=0.0001), and also was elevated in Liquidators sampled in 1991–1992 (72%, p=0.0076), 1993–1994 (22%, p=0.037), and 1995–1996 (62%, p=0.0001). In summary, HPRT MF was found to be the more sensitive and persistent indicator of radiation exposure, but the specificity of total gene deletions led to detection of probable heterogeneity of radiation exposure within the exposed population.  相似文献   

14.
Yield stress of 6% (w/w) waxy maize (WXM), cross-linked waxy maize (CLWM), and cold water swelling (CWS) starches in xanthan gum dispersions: 0%, 0.35%, 0.50%, 0.70%, and 1.0% was measured with the vane method at an apparent shear rate of 0.05 s−1. The intrinsic viscosity of the xanthan gum was determined to be: 112.3 dL/g in distilled water at 25 °C. Values of the static (σ0s) and dynamic (σ0d) yield stress of each dispersion were measured before and after breaking down its structure under continuous shear, respectively. The WXM and CWS starches exhibited synergistic behavior, whereas the CLWM starch showed antagonistic effect with xanthan gum. The difference (σ0s − σ0d) was the stress required to break the inter-particle bonding (σb). The contributions of the viscous (σv) and network (σn) components were estimated from an energy balance model. In general, values of σb of the starch–xanthan gum dispersions decreased and those of σn increased with increase in xanthan gum concentration.  相似文献   

15.
In this study, we have investigated the effect of urea on the physicochemical (pH and conductivity) and rheological properties of the chitosan-β-GP system in order to assess the main polymer–polymer interactions at low and high temperature. The pH of the solutions was slightly increased due to the consumption of H+ in solution through the hydrolysis of urea. Furthermore, the addition of urea considerably decreased the conductivity, and therefore the ionic strength of the solutions, and this effect was more important at high temperature. It indicated that urea strongly affects polymer–polymer interactions by weakening hydrogen bonds at low temperature, but in addition can hinder hydrophobic effect at high temperature since the reduction of ionic strength results in less screening of electrostatic repulsion between protonated glucosamine groups. At 15 °C, the addition of urea to chitosan-β-GP solutions decreased their elasticity, shortened relaxation times and simplified the relaxation process due to the disruption of hydrogen bonds. Heat-induced gelation of the chitosan system in the presence of urea showed higher gelation temperature (Tgel) in non-isothermal tests and longer gelation time (tgel) in isothermal conditions. The activation energy for gelation also increased with increasing urea concentration. We concluded that the detrimental effect of urea on the gelation process was mainly related to a decrease in polymer-polymer hydrophobic effect, as shown by the decrease in conductivity.  相似文献   

16.
Alkaline p-nitrophenylphosphate phosphatase (pNPPase) from the halophilic archaeon Halobacterium salinarum (previously halobium) was solubilized in reversed micelles of cetyltrimethylammonium bromide (CTAB) in cyclohexane with 1-butanol as cosurfactant. The hydrolysis reaction appears to follow Michaelis–Menten kinetics. The dependency of the maximum reaction rate (Vmax) on the water content θ (% v/v) (or ω0 value: molar ratio of water to surfactant concentrations) showed a bell-shaped curve for 0.3 M CTAB, but not for 0.2 M CTAB. The enzyme activity increased with the surfactant concentration at a constant ω0 value (10.27). When the surfactant concentration was increased at a constant θ, the enzyme activity decreased. The enzyme was more stable in reversed micelles than in aqueous media.  相似文献   

17.
Alpha-amylase of B. amyloliquefaciens was used for the structural characterization of the amylopectin from waxy-rice starch. Fractions of -dextrins with a degree of polymerization (d.p.) <5000 were isolated from amylopectin hydrolysates after 1 and 3 h. φ,β-Limit dextrins were prepared by successive phosphorolysis and beta-amylolysis of the fractions and these were analysed by a second alpha-amylolysis. Based on the hydrolysis pattern, the limit dextrins were divided into two major groups, A and B, which possessed units of clusters of d.p. 100–200 and 90–130, respectively. An extensive alpha-amylolysis resulted in characteristic distributions of dextrins with d.p. <80 which represented branched building blocks. Type A dextrins possessed more larger building blocks with d.p. 40, but less intermediate and small blocks, than type B. The φ,β-limit dextrin of the original amylopectin had a distinct distribution enriched in small building blocks. A model is proposed in which the two types of dextrins originate from regular and less regular structural domains of the amylopectin fraction within the starch granules.  相似文献   

18.
The seagrass Halophila decipiens Ostenfeld was grown axenically in a culture medium consisting of 20% artificial seawater, f/4 nutrients (except that glutamic acid was the nitrogen source), and 1% sucrose (w:v). The culture medium was adjusted to pH 5.0. A root–rhizome layer was created by solidifying a portion of the medium with 0.9% agar (w:v) and 1% activated charcoal (w:v). The rhizome layer also contained the following vitamins: 0.5 mg l−1 nicotinic acid, 0.5 mg l−1 pyridoxine, 0.5 mg l−1 biotin, 0.5 mg l−1 cyanocobalamin and 0.1 mg l−1 of thiamine HCl. The liquid overlay (without vitamins or charcoal) was poured onto the agar-solidified root–rhizome layer. Growth of H. decipiens was not improved by the addition of the auxins indoleacetic acid (IAA), indolebutyric acid (IBA) or naphthaleneacetic acid (NAA) at either of the tested concentrations (10 and 50 μM). At a concentration of 10 μM, the cytokinins 6-(γ,γ-dimethylallylamino) purine (2iP) and benzylaminopurine (BA) stimulated shoot and branch production compared to controls with no cytokinins. Among the tested nitrogen sources, growth was best on 1.7 mM glutamic acid. Cultures grown on 1.7 mM NH4Cl showed the same growth rates as those grown on glutamic acid, but the leaves were smaller and curled, suggesting incipient ammonium toxicity. Use of nitrate or urea led to mortality of the cultures. Long term axenic culture of H. decipiens appears to require the added vitamins. Hence, H. decipiens is the first seagrass known to need exogenous vitamins. Cultures of H. decipiens died when grown suspended in liquid cultures or in a biphasic medium system without activated charcoal in the root–rhizome layer. The use of more highly charged κ-carrageenan could not replace the use of activated charcoal and agar in the root–rhizome layer.  相似文献   

19.
Axin is encoded by the fused locus in mice and is required for normal vertebrate axis formation. It has recently been shown that axin associates with the adenomatous polyposis coli gene product (APC), β-catenin and glycogen synthase kinase-3 (GSK-3) in a complex that appears to regulate the level of cytoplasmic β-catenin. We have identified the Xenopus homologue of axin through its interaction with GSK-3β. Xenopus axin (Xaxin) is expressed maternally and throughout early development with a low level of ubiquitous expression. Xaxin also shows remarkably high expression in the anterior mesencephalon adjacent to the forebrain–midbrain boundary.  相似文献   

20.
A detailed study of the interaction of chitin molecular species with the solvent system N,N-dimethylacetamide (DMAc)/lithium chloride (LiCl) allowed the development of a new method for chitin fractionation by coacervate extraction. The controlled increase of the extracting power of the solvent was carried out using slight modification of the solvent composition. Partial extractions of molecular species were done between coacervation and complete dissolution limits using different mixtures of DMAc/LiCl of increasing extracting power. Fractions were characterized in DMAc/LiCl 5% (w/w) by viscometry and size exclusion chromatography with refractive index and multi-angle laser light scattering detectors. Fractions obtained by coacervate extraction range from 80,000 to 710,000 g mol−1 with polydispersity index between 1.28 and 1.44. The Mark–Houwink–Sakurada equation constants a and K for chitin in DMAc/LiCl 5% (w/w) were found to be 0.95 and 7.6×10−5 dl g−1, respectively.  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