首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The aim of the present study was to determine the influence of the ionic strength on the thermal stability of xanthan, i.e. xanthan resistance to chain breaking at high temperatures. Xanthan solutions of various ionic strengths were kept at 80, 90 and 95°C for periods up to 95 h. The thermal stability was determined by measuring the intrinsic viscosity after the heating periods. The experiments showed a critical ionic strength for the thermal stability of xanthan between 10 and 100 m NaCl or KCl in this temperature range. Below the critical ionic strength the intrinsic viscosity was rapidly reduced, whereas above the critical ionic strength the intrinsic viscosity was virtually unaffected by heating.

We then looked for a possible correlation between thermal stability and secondary structure of xanthan. The transition ionic strength (Im) of xanthan solutions, i.e. where xanthan is midway between an ordered and a disordered structure, was determined by NMR at constant temperatures. Im was found to be in the range of 24 m at 80°C to 60 m NaCl at 95°C, thus lying in the range of the critical ionic strength of the thermal stability. This suggests a close relationship between thermal stability and secondary structure of xanthan, indicated by the enhanced thermal stability in the ordered state. We believe this enhanced thermostability arises from a double-stranded conformation in the ordered state, as in DNA. The presence of double-stranded xanthan is also indicated by electron micrographs taken at both high and low ionic strengths.

The transition temperature (Tm) of xanthan was determined by NMR and optical rotation measurements. At the ionic strength of 7·5 m the two methods resulted in Tm values of 67 and 52°C respectively. This difference in Tm can possibly be due to the fact that the observed NMR and optical rotation (OR) effects are caused by different molecular phenomena.  相似文献   


2.
Aqueous NaCl solutions of dimerized Na xanthan with salt concentrations of 0.005, 0.01 and 0.1 were exposed to 80°C for different time periods t, and their viscosities were determined as a function of t. The measured relative viscosities decreased markedly with t, suggesting that Na xanthan denatured at 80°C undergoes some conformation changes or degradation. The molecular weights of the test samples recovered at different t were estimated by viscometry in cadoxen, a single-coil solvent for xanthan, and were found to decrease monotonically with t. Thus, it was concluded that the observed decreases in relative viscosity are due primarily to degradation of Na xanthan.  相似文献   

3.
Xanthan gum fermentation represents a good model for the study of the mixing of rheologically complex culture broths. Most of the previous work on power consumption dealt with ‘standard’, single impellers and used model fluids to simulate xanthan broths. This work describes the characterization of three dual-impeller combinations (D/T = 0·53) for the mixing of dehydrated—reconstituted fermentation broths of Xanthomonas campestris that had matched rheology to the actual broths. The bottom impeller was a Rushton turbine (RT) and the top impeller was another RT, a 45° pitched blade turbine (PT) or an A-310 Lightnin mixer (A310). The experiments were carried out in a tank of 0·0094 m3 working volume equipped with an air bearing dynamometer. The power was measured in a wide range of xanthan concentrations (5–40 kg m−3) in aerated (0·25, 0·5 and 1·0 vvm) and unaerated conditions. Unaerated power number (Po) vs. Reynolds number (Re) curves showed similar trends for the three combinations. Exponents close to −1 were obtained in the laminar region. A minimum in Po (Pomin) occurred at Re = 30–40, then increasing to a plateau value which was evident at Re> 200. In the transition region Pomin values were 4·3 (RT and RT), 3·6 (RT and PT) and 2·4 (RT and A310). The aerated power data for (RT and PT) and (RT and A-310) showed higher torque instabilities than the dual RT combinations at higher xanthan concentrations. The higher the xanthan concentrations, the higher the drop in power and the less important the effect of the aeration rate. Among the combinations tested, when using Rushton turbines, the well-mixed ‘cavern’ reached the tank wall (i.e., fluid motion was observed) at the lowest volumetric power input. High  相似文献   

4.
The viscosity in the low shear rate Newtonian domain of three biopolymers, locust bean gum, guar gum and xanthan gum was studied as a function of temperature and of polymer concentration in various aqueous solvents. The intrinsic viscosities [η]o of both galactomannans are not modified in the presence of 10 or 40% sucrose. In this case, a master curve relating the Newtonian specific viscosity (ηsp)o, to the reduced concentration c[η]o is obtained and allows (in good agreement with theoretical conjectures), two critical concentrations C* and C** to be defined, from which the value of the expansion coefficient may be estimated. For xanthan, as expected for a polyelectrolyte, [η]o depends strongly on salt concentration and on added sucrose and the results did not obey the above-mentioned master curve. However, it is shown that (ηsp)o depends only on xanthan concentration whenC > C**, and then it is assumed that chain dimensions have attained their unperturbed values whatever the solvent. Considering that both types of chains, random coils (galactomannans) and semi-rigid (xanthan) should give the same (ηsp)o-C[η]o master curve for C > C** when [η]o is replaced by its unperturbed counterpart [η]θ, a method for estimating [η]θ for the xanthan sample is proposed. In conclusion, the numerous exceptions to the widely accepted (ηsp)o vs C[η]o “universal” behaviour are mainly ascribed to significant differences in expansion coefficient values which depend on both the polymer and the solvent.  相似文献   

5.
The influence of xanthan conformation on the physicochemical behaviour of their mixtures with galactomannan from Schizolobium parahybae mannose:galactose ratio (M/G=3), was studied by viscoelastic measurements, differential scanning calorimetry (DSC) and chiroptical (circular dichroism) methods. The results suggested a more effective interaction of the galactomannan with disordered xanthan segments, which are more abundant in low salt concentrations but are still present in lower proportion at temperatures lower than the temperature of xanthan conformational transition (Tm). The dependence of ellipticity with temperature in a circular dichroism (CD) spectra suggested an ordering of the xanthan chains induced by galactomannan at the temperature of gel formation (Tg≈25°C), under conditions where xanthan alone exhibits a disordered conformation. The lower Tg value found (≈25°C) compared with that (60°C) usually described in the literature is certainly related to the M/G ratio and the galactosyl unit distribution along the mannan main chain.  相似文献   

6.
In-situ gelation of semidilute xanthan solutions with trivalent chromium, aluminum or iron ions was studied by rheology and UV-spectroscopy. Measurements of the elastic modulus of xanthan gel cylinders prepared by dialysis against the complexing ion at pH values from 2 to 6 indicate that monomeric species of the ion are ineffective, whereas dimeric or higher oligomeric species are effective in crosslinking the polysaccharide. When chromium was used as the crosslinking species, the dependence of the gelation rate on the ionic concentration followed a power law with a coefficient of 1·7. The gelation time and the gelation rate were found to extrapolate to zero at 1 m Cr for 2·5 mg/ml xanthan. The limiting concentration of xanthan needed for gelation with 5 m Cr(III) at 20°C was estimated as 0·35 mg/ml. This critical xanthan concentration is close to the overlap concentration c* estimated from the experimentally determined intrinsic viscosity [η] using c* = 1·4/[η]. An apparent activation energy for crosslinking of xanthan was calculated as Ea = 42 kJ/mol and Ea = 108 kJ/mol for Cr and Al ions, respectively. The fractal dimensionality of xanthan-Cr at the sol-gel transition was estimated as 1·3 applying the Chambon-Winter criterion for gelation, thus indicating that this gelation criterion is applicable also to stiff-chain polysaccharides such as xanthan.  相似文献   

7.
Flax pectins, with a low anhydrogalacturonic acid content (AGA≤50%), have been studied by viscometry, light scattering and size exclusion chromatography. After removal of interfering superstructures, intrinsic viscosities were correlated with weight average molecular weights to give a Mark-Houwink coefficient a=0·69. The persistence length of these heterogeneous chains (PI: q=20–25 Å), compared with that of homogalacturonan ones previously studied (PII–PIII: q=67 Å), shows a higher segment density (branched conformation). Using a wide range of experimental data ((η), Mi, Kd) obtained on narrow pectic fractions of known composition, specific calibration curves of the Sephacryl 200/0·2 NaCl/Flax pectins (PI or PII–PIII) system were established. Applied to the indirect characterization of flax pectins, this size exclusion chromatography (SEC) system constitutes a very convenient method for analysing the molecular and macromolecular parameters ((η), Mw, Mn, Ip, UV) or for controlling the molecular weight distribution of pectins in flax cell walls during the physiological steps of growth and retting.  相似文献   

8.
The aim of this work is to examine the possible interaction and extent thereof of the polysaccharide galactomannan (GAL) with the cereal protein gliadin (GLI) and a peptic-tryptic degraded gliadin (PT-GLI) by analytical ultracentrifugation. The work is part of a series of investigations into the field of coeliac disease (gluten-induced enteropathy) as gliadins are known to be toxic for patients with this disease.

The molecular integrity of the GAL and GLI preparations was first checked by sedimentation velocity and sedimentation equilibrium. Sedimentation velocity showed single boundaries indicating homogeneity and low-speed sedimentation equilibrium gave plausible apparent weight average molar masses of 180,000 g/ mol for GAL and 20,000 g/mol for GLI. PT-GLI, GLI and GAL in phosphate buffer (pH 6.5) and the incubated mixtures (stirred for 3 h at 37 °C; PT-GLI: GAL = 3.53:1, wt.wt.; GLI:GAL = 0.23 and 0.55:1, wt.wt.) were then investigated by sedimentation velocity at a temperature of 20 °C. The plots of 1/s20 vs. c of GAL, PT-GLI-GAL and GLI-GAL mixtures after incubation show a significantly different shape suggesting the presence of interactions. According to the equation 1/s20 = 1/so20(1 + ksc), values for {so20, ks} of {(4.02 ± (490.9 ± 28.9) ml/g, {(5.92 ± 0.24) S, (1152 ± 44) ml/g} and {(5.38 ± 0.19) S, (1141 ± 38) ml/g} for GAL and PT-GLI-GAL and GLI-GAL mixtures, respectively, were obtained. The concentration of GAL ranged from 0.75–3.0 mg/ml for GAL alone and from 0.34–1.50 mg/ml in the incubated mixtures. This apparent indication for a weak non-covalent protein-polysaccharide interaction was further supported by UV absorption spectrometry and gel filtration.  相似文献   


9.
The solution properties of hydroxyethyl cellulose (HEC) and hydrophobically modified hydroxyethyl cellulose (HM-HEC) have been investigated by means of viscometric and spectroscopic techniques involving free radical and fluorescent probes. The greater viscosity of HM-HEC solutions above a critical polymer concentration (Cp) of approximately 0·2% has been interpreted in terms of the formation of a three-dimensional network structure in which the polymer chains are effectively crosslinked by the intermolecular association of neighbouring hydrophobic side chains. Cp is considerably less than the predicted polymer coil overlap concentration (C*) of approximately 1%.

The interaction of the polymers with an anionic surfactant, sodium dodecyl sulphate (SDS) has also been investigated. A mechanism involving the interaction of free surfactant with the regions of intermolecular hydrophobic association is suggested to account for the considerable differences in the rheological behaviour of the polymers in the presence of SDS.  相似文献   


10.
Differential scanning calorimetry, high and small angle X-ray diffraction analyses have been carried out on air-dried and rehydrated rat tail tendon collagen in order to test the reversibility of collagen thermal denaturation. The mean enthalpy values calculated for the denaturation process of air-dried and rehydrated samples are ΔHD = 9.0 ± 0.8 cal/g and ΔHD = 11.9 ±0.7 cal/g respectively, while the denaturation temperatures are TD = 112 ± 1°C and TD = 51 ± 1°C. Partial reversibility of the coiled coil—random coil process can be obtained by storing the samples in air or more rapidly by equilibration in water. After denaturation air-dried collagen fibres recover not only their molecular structure but also their characteristic fibrillar structure. The latter does not greatly influence the mean experimental enthalpy values.  相似文献   

11.
A method of determining of the effective diffusion coefficient of substrate in a particle, where the diffusion and consumption of substrate by biocatalytic reaction are present simultaneously, was designed and experimentally verified. The method is based on measuring the overall rate of heterogeneous biocatalytic reaction in particles of varying diameter. The effective diffusion coefficient, De, was determined by fitting the measured reaction rates with the solution of the reaction-diffusion equation. The method is tailored for cases where the enzyme reaction is governed by the Michaelis-Menten kinetics. The value of Km required for the solution of the mathematical model was adopted from the measurement of the kinetics of free cells, whereas the rate parameter, k2, was optimized together with De. As an experimental model, the sucrose hydrolysis catalyzed by Ca-alginate-entrapped yeast cells was examined. The particle diameter varied in the range of 1.2–3.9 mm and the initial reaction rates were measured in a batch-stirred reactor at a sucrose concentration of 100 m . The De of sucrose at 30°C was found to be 2.9 · 10−10 m2s−1.  相似文献   

12.
The formation of three [Tl(en)n]3+ complexes (n=1–3) in a pyridine solvent has been established by means of 205Tl and 1H NMR. Their stepwise stability constants based on concentrations, Kn=[Tl(en)n 3+]/{[Tl(en)n−1 3+]·[en]}, at 298 K in 0.5 M NaClO4 ionic medium in pyridine, were calculated from 205Tl NMR integrals: log K1=7.6±0.7; log K2=5.2±0.5 and log K3=2.64±0.05. Linear correlation between both the 205Tl NMR shifts and spin–spin coupling 205Tl–1H versus the stability constants has been found and discussed. A single crystal with the composition [Tl(en)3](ClO4)3 was synthesized and its structure determined by X-ray diffraction. The Tl3+ ion is coordinated by three ethylenediamine ligands via six N-donor atoms in a distorted octahedral fashion.  相似文献   

13.
The effect of acidification on a typical commercial xanthan and on pyruvate-free xanthan (PFX), alone and in gelling mixtures with konjac glucomannan (KGM), has been studied by differential scanning calorimetry (DSC) and small-deformation oscillatory measurements of storage modulus (G′) and loss modulus (G″). For both xanthan samples, progressive reduction in pH caused a progressive increase in temperature of the disorder–order transition in DSC, and a progressive reduction in gelation temperature with KGM. This inverse correlation is interpreted as showing that synergistic gelation involves disruption of the xanthan 5-fold helix, probably by attachment of KGM to the cellulosic backbone of the xanthan molecule (as proposed previously by a research group in the Institute of Food Research, Norwich, UK). Higher transition temperature accompanied by lower gelation temperature for PFX in comparison with commercial xanthan at neutral pH is explained in the same way. However, an additional postulate from the Norwich group, that attachment of KGM (or galactomannans) can occur only when the xanthan molecule is disordered, is inconsistent with the observation that gelation of acidified mixtures of KGM with PFX can occur at temperatures more than 60 °C below completion of conformational ordering of the PFX component (as characterised by DSC). Increase in G′ on cooling for mixtures of commercial xanthan with KGM at pH values of 4.5 and 4.25 occurred in two discrete steps, the first following the temperature-course observed for the same mixtures at neutral pH and the second occurring over the lower temperatures observed for mixtures of KGM with PFX at the same values of pH. These two “waves” of gel formation are attributed to interaction of KGM with, respectively, xanthan sequences that had retained a high content of pyruvate substituents, and sequences depleted in pyruvate by acid hydrolysis. At pH values of 4.0 and lower, gelation of mixtures of KGM with commercial xanthan followed essentially the same temperature-course as for mixtures with PFX, indicating extensive loss of pyruvate under these more strongly acidic conditions. Mixtures prepared at pH values in the range 4.0–3.5 gave comparable moduli at room temperature (20 °C) to those obtained at neutral pH, but showed substantial softening on heating to body temperature, suggesting possible applications in replacement of gelatin in products where “melt-in-the-mouth” characteristics are important for acceptability to the consumer.  相似文献   

14.
A xanthan sample with molecular weight M = 2.2 x 10(6) was investigated in three solvents: bidistilled water, 0.2 M aqueous NaCl and cadoxen by flow birefringence and viscometry methods in dilute solutions. It was shown that the optical shear rate coefficients of xanthan in aqueous and cadoxen media differ by two orders of magnitude. An estimation of xanthan optical anisotropy in different conformational states has been made and compared with values for other polysaccharides: dextran, pullulan, cellulose and chitosan. The process of denaturation and the flow birefringence of renaturated xanthan in aqueous solutions (after heat treatment at 121 degrees C) have also been studied.  相似文献   

15.
Association of bovine serum albumin (BSA) on heating in the presence and absence of 2% xylose has been studied using dynamic light scattering and sedimentation velocity. When 3% solutions of the protein alone are heated at 95°C association products are formed with molar masses of 2 × 106g/mol, a value which is independent of the time of heating. These aggregates can be dissociated in solvents that disrupt non-covalent bonds. When the reducing sugar xylose is present there is a continuous change in the hydrodynamic properties with time. After 80 min a molar mass in excess of 7 × 106g/mol is obtained. This increase in molar mass is attributed to additional non-disulphide linkages resulting from the Maillard reaction. Information about the gross conformation of the Maillard induced association products has been obtained from MHKS (Mark-Houwink-Kuhn-Sakarada) double logarithmic plots of D20,w and s20,w against molar mass. The values of the MHKS coefficients obtained are most consistent with a linear rod: i.e. the association is of an end-to-end type  相似文献   

16.
Aqueous solutions of fractions of an extracellular linear mannan formed by Rhodotorula rubra yeast have been investigated by hydrodynamic methods (high-speed sedimentation, translation isothermic diffusion and viscometry). The molecular weight was determined according to Svedberg ( ) and the polydispersity parameters of the initial sample were also determined (Mw/Mn = 1·20 and Mz/Mw = 1·21). Relationships between the molecular weight (M) and so, Do and [η] in the range were: [η] = 2·33 × 10−2 M0.75, Do = 1·65 × 10−4 M0·58, so = 2·24 × 10−15 M0·43. The equilibrium rigidity and hydrodynamic diameter of chains representing mannan molecules were evaluated.  相似文献   

17.
The oscillatory rheology of one rodlike and one semiflexible xanthan sample has been investigated as a function of temperature in the range of xanthan concentrations where the polymer forms a lyotropic liquid crystalline phase in aqueous NaCl solutions. Readily observed changes in the rheological observables at temperatures corresponding to phase boundaries permit construction of the biphasic chimney region of the temperature-composition phase diagram. The chimney region leans toward larger values of the polymer concentration with increasing temperature, presumably as a consequence of a reduction in the effective axial ratio of the helical polymer with increasing temperature. The results permit construction of plots of the rheological observables as a function of polymer concentration at temperatures T in the range 20 相似文献   

18.
Kinetic study of anaerobic digestion of brewery wastewater   总被引:2,自引:0,他引:2  
A study of the kinetics of the anaerobic digestion of brewery wastewater was carried out using a 1-litre, continuous-flow, completely-mixed, bioreactor operating at 35°C and containing a saponite-immobilized biomass at a concentration of 6·2 g volatile suspended solids (VSS)/litre. The bioreactor worked satisfactorily in a range of hydraulic retention times from 1·2 to 10 days and eliminated more than 95% of the initial chemical oxygen demand (COD) in all instances.

Guiot's kinetic model was used to determine the macroenergetic parameters of the system, and showed it to have a yield coefficient for the biomass (Y) of 0·080 g VSS/g COD and a specific rate of substrate uptake for cell maintenance (m) of 0·045 g COD/g VSS day.

The experimental results showed the rate of substrate uptake (Rs; g COD/g VSS day), correlated with the concentration of biodegradable substrate (Sb; g COD/litre), through an equation of the Michaelis-Menten type.  相似文献   


19.
Six samples of Na xanthan in 0.01M aqueous NaCl containing 0.01 N HCl (pH = 2) were studied by light scattering and viscosity. This study was motivated by the finding that the intrinsic viscosity [η] fairly sharply decreased when the pH of the solvent was lowered from about 6 to 2 by adding HCl to 0.01M aqueous NaCl in which Na xanthan dissolves as rigid dimers having a double-helical structure. The data for weight-average molecular weight, radius of gyration, and [η] showed that Na xanthan at pH = 2 remains a dimer behaving as a semiflexible chain. Data analysis in terms of known theories for unperturbed wormlike chains yielded 0.47 ± 0.02, 2.0 ± 0.6, and 68 ± 7 nm for the contour length h per main-chain residue, diameter d, and persistence length q of the dimer, respectively. these h and d values agreed with the pitch per main-chain residue and the diameter of the double helix of Na xanthan in 0.01 or 0.1M aqueous NaCl. However, the q value, which was close to the intrinsic persistence length q0 ( = q in the absence of electrostatic interaction) of Na xanthan at pH = 2, was much smaller than the q0 (106 nm) of this helix. We concluded that the xanthan dimer at pH = 2 assumes a double-helical structure, which is geometrically the same as, but is more flexible than, that at neutral pH.  相似文献   

20.
The lack of a truly satisfactory sensor which can characterize the thermal environment at the spatial scale experienced by small endotherms has hindered study of their thermoregulatory behavior. We describe a general design for a rugged, easily constructed sensor to measure standard operative temperature, Tes. We present specific designs for adult dark-eyed juncos (Junco hyemalis) and hatchling mallards (Anas platyrhynchos). Sensor response was stable and repeatable (±1.4%) over the course of several months. Over the range of conditions for which validation data were available (variable air temperature and wind with negligible net radiation), sensors predicted the mean net heat production of live animals to within ±0.023 W (equivalent to ±1°C at Tes=15°C). The main limit on accuracy was scatter in the data on metabolism and evaporative water loss in live animals. These sensors are far more rugged and easily constructed than the heated taxidermic mounts previously used to measure Tes. These characteristics facilitate the use of significant numbers of sensors in thermal mapping studies of endotherms.  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