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.
Xanthan gum and sodium caseinate are used to improve stability and texture of food. To investigate interactions between them, the effects of pH on structure of sodium caseinate–xanthan gum complex were analyzed. HCl titration showed that the absorbance of the mixture was different from that of sodium caseinate alone throughout the acidification, and that syneresis in the mixture was delayed in acidic pH. Rennet digestion clarified that xanthan gum retarded degradation of κ-casein at pH 2.7. Atomic force microscopy revealed that xanthan gum interaction with sodium caseinate was pH-dependent. Sodium caseinate particles were individually bound with xanthan gum at pH 6.6, and a side-by-side aggregation of sodium caseinate along xanthan gum was observed at pH 4.2. The mixture formed a network composed of rod-like fibers at pH 2.7. These results indicate that hydrophobic and electrostatic interactions play a role in the complex formation at neutral and acidic pH, respectively.  相似文献   

2.
A membrane-covered polarographic oxygen electrode was used to measure oxygen diffusion coefficients in aqueous polyelectrolyte solutions of xanthan gum, sodium alginate, and sodium carboxymethylcellulose (CMC). In sodium alginate solutions, dilute xanthan solutions, and solutions containing more than 0.3 wt % CMC, oxygen diffusion coefficients decrease with increasing polymer concentrations. Interestingly, in dilute CMC solutions and concentrate xanthan solutions containing more than 0.5 wt % xanthan gum, oxygen diffusion coefficients increase with increasing polymer concentrations, and values exceeding that in pure water are generally observed.  相似文献   

3.
The non-Newtonian behavior and dynamic viscoelasticity of solutions of the Ca salt of xanthan were measured with a rheogoniometer. The Ca salt of xanthan showed pseudoplastic behavior < 0.1 % but was plastic >0.3%. Compared with native, Na, and K salts of xanthan, the Ca salt had higher dynamic viscoelasticity at high concentrations. The apparent viscosity of Ca salt of xanthan was very large at low temperatures and decreased with increasing temperature. These suggest very strong intermolecular association of the xanthan (Ca salt) molecules, probably due to the formation of ionic force between adjacent charged trisaccharide side-chains via Ca2+ on different molecules. Possible structures for quaternary association including intramolecular association of xanthan molecules (Ca and K salts) in aqueous solution were proposed.  相似文献   

4.
Ye A  Flanagan J  Singh H 《Biopolymers》2006,82(2):121-133
The formation of electrostatic complexes between sodium caseinate and gum arabic (GA) was studied as a function of pH (2.0-7.0), using slow acidification in situ with glucono-delta-lactone (GDL) or titration with HCl. The colloidal behavior of the complexes under specific conditions was investigated using absorbance measurements (at 515 or 810 nm) and dynamic light scattering (DLS). In contrast to the sudden increase in absorbance and subsequent precipitation of sodium caseinate solutions at pH < 5.4, the absorbance values of mixtures of sodium caseinate and GA increased to a level that was dependent on GA concentration at pH 5.4 (pH(c)). The absorbance values remained constant with further decreases in pH until a sudden increase in absorbance was observed (at pH(phi)). The pH(phi) was also dependent upon the GA concentration. Dynamic light scattering (DLS) data showed that the sizes of the particles formed by the complexation of sodium caseinate and GA between pH(c) and pH(phi) were between 100 and 150 nm and these nanoparticles were visualized using negative staining transmission electron microscopy (TEM). Below pH(phi), the nanoparticles associated to form larger particles, causing phase separation. zeta-Potential measurements of the nanoparticles and chemical analysis after phase separation showed that phase separation was a consequence of charge neutralization. The formation of complexes between sodium caseinate and GA was inhibited at high ionic strength (>50 mM NaCl). It is postulated that the structure of the nanoparticles comprises an aggregated caseinate core, protected from further aggregation by steric repulsion of one, or more, electrostatically attached GA molecules.  相似文献   

5.
Quasielastic light-scattering and other physical-chemical techniques have been used to compare the conformation and intermolecular interactions of xanthan in water, aqueous sodium chloride, and urea solutions. The results showed that xanthan dissolved in 4m urea has a disordered conformation after the solution has been maintained for 3 h at 95° and then cooled to room temperature. This conformation is similar to that previously observed only in solutions having low ionic strength at higher temperatures, following disruption of the ordered, low-temperature form. “Anomalous” behavior is seen for xanthan as a function of ionic strength, in that the hydrodynamic radius increases with increase in ionic strength, whereas a decrease is typical for polyelectrolytes. These observations suggest that aggregation of rod-like chains, similar to that seen for other stiff-chain polymers, occurs for xanthan in salt solutions, where the charged groups of the polyelectrolyte are screened by the salt ions. This aggregation may explain some of the high values reported in the literature for the molecular weight.  相似文献   

6.
The differences in response of 1% potato and 4% maize starch pastes to sodium caseinate inclusion were investigated. Pasting of the starches was performed at 95 °C for l h in a range of concentrations of sodium caseinate. Caseinate levels as low as 0.01% dramatically reduced the swelling volume of potato starch and hence the viscosity of the system. Since sodium chloride addition shows similar effects, it appears that caseinate acts through a non-specific ionic strength effect. The influence of caseinate on maize starch was less clear since it depended on the solvent medium. In distilled, deionized water, there was an increase in viscosity with increasing caseinate concentration, which may simply be explained by a contribution of the caseinate to the viscosity of the continuous phase. However, in 0.1M, pH 7.0 buffer the results suggest that caseinate may inhibit retrogradation as the viscosity of the system after ageing is reduced by its inclusion. It is suggested that phase separation between starch and caseinate is encouraged at high salt concentrations. As a consequence, both starch granule swelling and subsequent retrogradation are discouraged by caseinate in the buffer system, but not when pasting is carried out in distilled, deionized water.  相似文献   

7.
1. The deposition of casein on a platinum anode which takes place on the passage of a direct current through solutions of alkali caseinates was quantitatively studied, and it was found that: (a) the amount of casein which is deposited is directly proportional to the current, i.e. it obeys Faraday''s law; (b) the amount of casein deposited is inversely proportional (within the limits studied) to the amount of alkali which is combined with the casein. 2. A method of determining the transport numbers of proteins insoluble at their isoelectric point has been developed. 3. A titration method for determining the amount of alkali in a casein solution is given. 4. Data from the results of transference experiments on sodium caseinate, potassium caseinate, cesium caseinate, and rubidium caseinate solutions are given. It is shown that the data are best explained on the assumption that in these solutions the carriers of the current are alkali metal cations and casein anions. 5. On the basis of our transference results an explanation is given of the results which were obtained by Robertson and by Haas in their migration experiments.  相似文献   

8.
Properties of agents that effectively entrap liquid lipids.   总被引:1,自引:0,他引:1  
A droplet of an oil-in-water emulsion of methyl linoleate in a saccharide or protein solution that contained with a surfactant, a stabilizer, or both was dehydrated by drying equipment for a single droplet that resembled a spray drier. The lipid exposed on the surface of dehydated samples was extracted and measured by gas chromatography. Gum arabic or gelatin without additives resulted in little lipid being exposed; they were good entrapping agents. Little lipid was exposed with a pullulan solution containing lecithin, sugar ester, carboxymethylcellulose, or sodium caseinate but much was exposed with a maltodextrin solution containing any of the surfactants tested. When both the surfactant lecithin and the stabilizer xanthan gum were added to the emulsion prepared in a maltodextrin solution, lipid was not detected. The results suggested that effective entrapping agents of liquid lipids cause much emulsification, stabilize the emulsion (that is, they cause the continuous phase to be very viscous), and create a dehydrated matrix of fine, dense network layers.  相似文献   

9.
The ionotropic gelation of double-layered emulsions composed of sodium caseinate and ??-carrageenan at pH values of 7 and 3.5 was evaluated, in order to obtain potential encapsulation matrices for hydrophobic compounds. The influence of some of the extrusion process variables (nozzle diameter at fluid exit and collecting distance) on the microbead production was studied, as well as the stability of the microbeads. The fluid nozzle diameter showed little influence on the shape of the microbeads, with a slight tendency for a decrease in microbead diameter with increase in fluid nozzle diameter. On the other hand, the collecting distance strongly influenced the microbead shape and they became more spherical (aspect ratio was reduced from ~2.0 to ~1.4) as the collecting distance was increased from 10?cm to 50?cm. The emulsion pH did not affect the aspect ratio of the microbeads, but the diameter was greater for microbeads produced at pH 3.5. This difference was attributed to the kind of interactions occurring between the ??-carrageenan and sodium caseinate at these distinct pH values. The microbeads were highly unstable when dispersed in deionized water, sugar solutions and low salt concentrations, releasing the encapsulated oil. However, no release of oil from the microbeads was observed when they were dispersed in ethanol or potassium chloride solutions with concentrations above 0.75?%, although their shape was modified when dispersed in ethanol. In general, the results obtained demonstrated the viability of the extrusion process to produce biopolymer-based microbeads and the potential application of these systems.  相似文献   

10.
We investigated the effects of Tween emulsifier fatty acid chain length on the shear stability and crystallization behavior of 35 wt% partially crystalline oil-in-water emulsions prepared with and without 1 wt% sodium caseinate. Emulsions containing sodium caseinate and Tween 20, 40, 60 or 80 varied in shear stability, degree of supercooling and crystallization behavior depending on the type and concentration of Tween as well as the presence of protein. Generally, emulsions containing the unsaturated emulsifier Tween 80 were the most shear sensitive followed by the saturated emulsifiers Tween 20, 40 and 60 in order of increasing fatty acid chain length. Long chain saturated Tween emulsifiers (40 and 60) improved shear stability regardless of whether sodium caseinate was present indicating that alone, these emulsifiers form more robust interfacial films compared to the saturated short chain length Tween 20 and Tween 80. In emulsions prepared with sodium caseinate, the degree of supercooling decreased and the crystallization rate diminished with increasing saturated fatty acid chain length but only negligible changes were found without sodium caseinate. Together, these findings indicate that long chain saturated Tween emulsifiers provide better emulsion stability regardless of the presence of sodium caseinate but with sodium caseinate, stability may also be affected by changes to fat crystallization. These novel findings provide guidance on how combinations of proteins and emulsifiers can be used to modify and control the stability of partially crystalline oil-in-water emulsions through their combined effects on the properties of the interfacial film and fat crystallization.  相似文献   

11.
The influence of thermal treatment and operational conditions (pH and stirrer speed) used in the process of xanthan production by Xanthomonas arboricola pv pruni strain 106 were evaluated through yield of xanthan, aqueous solution and fermentation broth viscosity, sodium content, pyruvate and acetyl content and molar mass. Different conditions used during the fermentation affected the xanthan characteristics. Thermal treatment decreased the final yield and pyruvate and acetyl content, and increased the xanthan aqueous solution and fermentation broth viscosities, as well as molar mass. In this study the best combination of yield and viscosity was obtained with the use of pH 7 and 400 rpm during fermentation and post-fermentation thermal treatment. Aggregation of xanthan molecules promoted by heating and detected through an increase of molar mass was apparently affected by the sodium content. As a result, a correlation between molar mass and xanthan solution viscosity could be observed.  相似文献   

12.
We present findings on the influence of interfacial layer composition on the colloidal interactions and associated structural and material properties of oil-in-protein gel emulsions, as applied to a model Mozzarella cheese analogue. Model cheese samples were produced through thermal mixing of pre-prepared oil-in-water emulsions with a renneted casein gel. Sodium caseinate and Tween 20 were used as the emulsifiers. Microstructural analysis showed sodium caseinate stabilised droplets to be homogeneously dispersed within the cheese structure, whilst droplets stabilised by Tween 20 were phase concentrated into localised fat domains within the continuous protein network. Particle size measurements determined that, on chilled storage, the droplets in these localised regions underwent extensive partial coalescence, whilst the homogenously distributed caseinate droplets showed little change in droplet size. Small deformation rheology (4 to 80 °C) determined the sodium caseinate emulsion as providing a reinforcing effect on the protein network across the entire temperature range, while the Tween 20 emulsion was observed to mechanically strengthen the cheese structure at only at temperatures for which the fat phase was solid whilst serving to weaken the structure on transitioning to a molten state. Differences in droplet structure and stability were determined as influencing cheese melt and flow characteristics. During melting, no oiling-off observed for cheese samples comprising sodium caseinate stabilised droplets, compared to Tween 20 stabilised emulsions where extensive oiling-off was observed. Findings corroborate the hypothesis that caseinate coated droplets behave as active fillers within the protein network, whilst the Tween 20 stabilised emulsion are non-interactive.  相似文献   

13.
Optical rotatory dispersion (ORD) data were obtained for a sample of the sodium salt of xanthan gum dissolved in water-cadoxen mixtures at 25°C. The double-helical dimer of the polysaccharide was previously found to dissociate directly to single chains when wcad (the weight fraction of cadoxen in the mixed solvent) increases from 0·3 to 0·8. ORD data were also obtained for solutions prepared by diluting xathan solutions at low concentration in cadoxen with water to different wcad values. From previous work these were taken as those for single dissociated chains in the mixed solvent. These sets of data led to the finding that the specific rotation at 300 nm wavelength does not reflect the dissociation of the xanthan double helix in water-cadoxen mixtures, but the Moffitt parameter does. This parameter gave evidence that single coiling chains in cadoxen become intramolecularly ordered to a conformation similar to that of the individual chains in the double helix when the solution is diluted with water to wcad below 0·4.  相似文献   

14.
Proteolytic sporeforming bacteria capable of surviving processing heat treatments in synthetic or fabricated protein foods exhibited no antagonistic effects on growth of Clostridium perfringens, but instead shortened the lag of subsequent growth of C. perfringens in sodium caseinate and isolated soy protein. Bacillus subtilis A cells were cultured in 3% sodium caseinate or isolated soy protein solutions. The subsequent effect on the lag time and growth of C. perfringens type A (strain S40) at 45 C was measured by colony count or absorbance at 650 nm, or both. B. subtilis incubation for 12 h or more in sodium caseinate reduced the C. perfringens lag by 3 h. Incubation of 8 h or more in isolated soy protein reduced the lag time by 1.5 h. Molecular sieving of the B. subtilis-treated sodium caseinate revealed that all molecular sizes yielded a similar reduced lag time. Diethylaminoethyl-Sephadex ion exchange fractionation and subsequent amino acid analysis indicated that the lag time reduction caused by B. subtilis incubation was not related to charge of the peptides nor to their amino acid composition. Apparently the shortened C. perfringens lag in these B. subtilis-hydrolyzed food proteins was a result of the protein being more readily available for utilization by C. perfringens.  相似文献   

15.
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   

16.
Rheological measurements of dense calcium caseinate and sodium caseinate dispersions (> or =15%) provided insight into the factors determining shear-induced structure formation in caseinates. Calcium caseinate at a sufficiently high concentration (30%) was shown to form highly anisotropic structures during shearing and concurrent enzymatic cross-linking. In contrast, sodium caseinate formed isotropic structures using similar processing conditions. The main difference between the two types of caseinates is the counterion present, and as a consequence, the size of structural elements and their interactions. The rheological behavior of calcium caseinate and sodium caseinate reflected these differences, yielding non-monotonic and shear thinning flow behavior for calcium caseinate whereas sodium caseinate behaved only slightly shear thinning. It appears that the intrinsic properties of the dense caseinate dispersions, which are reflected in their rheological behavior, affect the structure formation that was found after applying shear. Therefore, rheological measurements are useful to obtain an indication of the structure formation potential of caseinate dispersions.  相似文献   

17.
A two‐part study determined the effectiveness of gum‐based rinses with or without oil for alleviating residuals of a bitter (0.8 g/L caffeine solution) and an astringent (1g/L alum solution) stimuli in serial responses using a sip and spit method. In Experiment 1, rinsing with deionized water was compared to rinsing with 0.3% xanthan gum in water alone or with 5% corn oil and 0.55% carboxymethyl cellulose (CMC) in water alone or with 5% or 10% corn oil. The 0.3% xanthan gum and 5% corn oil mixture resulted in the highest stimulus intensity difference before and after rinsing, whereas 0.55% CMC had a value of nearly zero for residual effect. Results of a two‐alternative forced choice test used in Experiment 2 indicated that both were equally effective for reducing bitterness residuals, but 0.55% CMC solution resulted in the lowest (p=0.007) residual effect for astringency. Therefore, the 0.55% CMC rinse was considered an effective interstimulus rinse to use for assessing both bitterness and astringency in model solutions.  相似文献   

18.
The structure and conformation of xanthan in aqueous solution following various processing treatments typically encountered in its application were investigated in this study. Treatments such as heating, autoclaving, high pressure homogenisation and irradiation were subjected to the same sample. Parameters such as weight average molecular weight (M(w)), polydispersity index, root mean square radius of gyration, intrinsic viscosity and Huggins constant were used to monitor the effect of these treatments. Additionally, we have quantified the mass recovery of samples examined by gel permeation chromatography and light scattering to properly account for all fractions present in xanthan solutions. Atomic force microscopy (AFM) images together with height measurements confirmed that xanthan conformation is double helical ordered renatured state (pre-heat treated by the manufacturer) in dilute solution conditions and random coil conformation in very dilute solution. The ordered (renatured) conformation is shown to have partially molten double helix, with more flexibility than the perfectly ordered native double helix. Heat treatment for 2h at 85°C reduces the M(w) of xanthan to half its initial value, and mass recovery measurements indicate that it completely overcomes its associative nature. Thermally treated xanthan solution in the dilute region leads to an order-disorder transition, as determined by contour length per unit mass. Similarly, irradiation of xanthan solution results in an order-disorder transition together with the production of single strand low molecular weight molecules. Autoclaving and high pressure homogenisation treatments cause degradation of xanthan. The results from treated xanthan solutions following high pressure homogenisation and irradiation confirm that xanthan does not reassociate. A revised summary of xanthan conformation in solution together with schematic models following the various treatments are proposed.  相似文献   

19.
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   

20.
Claviceps purpurea PRL 1980 grew on partially dissociated succinic acid (pH 4) but not on fully dissociated succinic acid (pH 7.2). Myeclium suspended in 42 mM solution of partially ionized succinic acid (pH 4; 60.1% nonionized, 39% monoanion, and 0.9% dianion, K+ salt) over a period of 25 min accumulated more succinic acid carbon than mycelium suspended in highly ionized solution (pH 6.8; 0.01% nonionized, 4.8% monoanion, and 95% dianion). The greater accumulation from partially ionized solution was not attributable solely to metabolism of succinic acid nor to the lower external concentration of potassium ion. Rate of uptake by sodium azide and iodoacetate-treated mycelium was proportional to external concentration at least up to 200 mumol/ml. External potassium or sodium ion was not required for uptake by inhibited or uninhibited mycelium and external sodium ion and glucose did not allow concentration of succinic acid. The internal concentrations of succinic acid carbon expressed as succinic acid in cell water were about the same as the external concentrations. Uptake was not appreciably affected by extent of ionization of external succinic acid but accumulation was markedly affected. A plot of accumulated succinic acid carbon against external pH produced a bimodal curve with the two maxima corresponding to the maximal concentrations of nonionized and monoanion succinic acid. The bimodal curve probably results from overlapping of two separate curves; the nonionized form accumulating efficiently because of one interaction with the cell and the monoanion form accumulating efficiently because of another interaction. Uptake from concentrated solution is by diffusion and efflux is rapid but not complete. Efflux is not retarded by presence of phosphate in the external solution.  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