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.
Glass transition temperatures of cassava starch (CS)-whey protein concentrate (WPC) blends were determined by means of differential scanning calorimetry (DSC) in a water content range of 8-20% (dry basis, d.b.). Water equilibration in the samples was carried out by storing them at room temperature (25 °C) during four weeks. Physical aging and phase segregation were observed in some samples after this storage period depending on the water content. Both, first DSC heating scans and tan δ curves of CS-WPC blends with intermediate water content (10-18%), showed two endothermic thermal events. The first one appeared at around 60 °C and was independent of water content. The second one was detected at higher temperatures and moved towards the low-temperature peak as the water content increased. The results can be explained by a phase segregation process that can take place when the samples are conditioned below their glass transition temperatures. The Gordon-Taylor equation described well the plasticizing effect of water on the blends. WPC was also found to decrease the glass transition temperature, at constant water content, an effect attributed to additional water produced during browning reactions in the blends.  相似文献   

2.
A novel biodegradable graft copolymer, dextran-g-poly(1,4-dioxan-2-one) (PODEX), was synthesized through the ring-opening polymerization (ROP) of 1,4-dioxan-2-one (PDO) onto a dextran backbone. Initially, dextran was silylated with 1,1,1,3,3,3-hexamethyldisilazane. The grafting from pathway was conducted with various (30–70) PDO/OH feed ratios to obtain PODEX copolymers with a various PPDO graft structures. Graft copolymers were characterized by FT-IR, 1H and 13C NMR, DSC, TGA, and WAXD. Molecular weights of the PODEX copolymers (MW: 94,700–117, 300 Da), glass transition temperature (−29 to −17 °C), melting temperature (82–100 °C), and crystallinity (32–40%) were increased with the content of PPDO. AFM observations revealed that polymeric micelles were spherical and uniform in morphology with around 30–58 nm diameter. Critical micelle concentration (CMC) of self-assembled system was significantly decreased from 3.2 to 1.09 mg/L with the increment of PPDO.  相似文献   

3.
The effect of water content on the glass transition temperatures of a ready to eat cereal formulation was determined, as well as for its major components, oat flour, rice flour and an oat–rice flour blend, in the same ratio as they are present in the formulation. All samples were compression moulded at high temperature and were moisture conditioned in a 10–22% interval (dry basis). Glass transition temperatures (Tg) were measured by differential scanning calorimetry (DSC) and the main mechanical relaxation temperatures (Tα), measured by dynamic mechanical thermal analysis (DMTA). The relaxation temperatures taken at tan δ peaks, were found 20–30 °C larger than Tg. Besides the plasticizing effect of water adequately described by the Gordon–Taylor equation, no differences of Tg (and Tα) values between the major components were obtained at a constant moisture content. The Tg and Tα values of the RTE formulation were found to be about 30 °C lower than its components, a result which was attributed to the plasticizing effect of the minor components in the formulation (sugar and malt extract).  相似文献   

4.
An improved method is described for the production of chitosan from mycelia of the fungus Gongronella butleri, grown by solid substrate fermentation on sweet potato. The chitosan was extracted subsequently by 11 M NaOH at 45 °C, and 0.35 M acetic acid at 95 °C. The resulting extract was clarified using a heat-stable, commercial -amylase. The yield (4–6 g/100 g mycelia) and relative number average molecular weight (44–54 kDa) of the chitosan increased with increasing duration of fungal growth up to the sixth day.  相似文献   

5.
Highly oriented fibers of Li-, Na-, K-, and CsDNA were prepared with a previously developed wet spinning method. The procedure gave a large number of equivalent fiber bundle samples (reference length, L0, typically = 12–15 cm) for systematic measurements of the fiber length L in ethanol–water solutions, using a simple mechanochemical set up. The decrease in relative length L/L0 with increasing ethanol concentration at room temperature gave evidence for the B-A transition centered at 76% (v/v) ethanol for NaDNA fibers and at 80 and 84% ethanol for K- and CsDNA fibers. A smaller decrease in L/L0 of LiDNA fibers was attributed to the B-C transition centered at 80% ethanol. In a second type of experiment with DNA fibers in ethanol–water solutions, the heat-induced helix–coil transition, or melting, revealed itself in a marked contraction of the DNA fibers. The melting temperature Tm, decreased linearly with increasing ethanol concentration for fibers in the B-DNA ethanol concentration region. In the B-A transition region, Na- and KDNA fibers showed a local maximum in Tm. On further increase of the ethanol concentration, the A-DXA region followed with an even steeper linear decrease in Tm. The dependence on the identity of the counterion is discussed with reference to the model for groove binding of cations in B-DNA developed by Skuratovskii and co-workers and to the results from Raman studies of the interhelical bonds in A-DNA performed by Lindsay and co-workers. An attempt to apply the theory of Chogovadze and Frank-Kamenetskii on DNA melting in the B-A transition region to the curves failed. However, for Na- and KDNA the Tm dependence in and around the A-B transition region could be expressed as a weighted mean value of Tm of A- and B-DNA. On further increase of the ethanol concentration, above 84% ethanol for LiDNA and above about 90% ethanol for Na-, K-, and CsDNA, a drastic change occurred. Tm increased and a few percentages higher ethanol concentrations were found to stabilize the DNA fibers so that they did not melt at all, not even at the upper temperature limit of the experiments (~ 80°C). This is interpreted as being due to the strong aggregation induced by these high ethanol concentrations and to the formation of P-DNA. Many features of the results are compatible with the counterion–water affinity model. In another series of measurements, Tm of DNA fibers in 75% ethanol was measured at various salt concentrations. No salt effect was observed (with the exception of LiDNA at low salt concentrations). This result is supported by calculations within the Poisson–Boltzmann cylindrical cell model. © 1994 John Wiley & Sons, Inc.  相似文献   

6.
Long-term storage of desiccated nucleated mammalian cells at ambient temperature may be accomplished in a stable glassy state, which can be achieved by removal of water from the biological sample in the presence of glass-forming agents including trehalose. The stability of the glass may be compromised due to a nonuniform distribution of residual water and trehalose within and around the desiccated cells. Thus, quantification of water and trehalose contents at the single-cell level is critical for predicting the glass formation and stability for dry storage. Using Raman microspectroscopy, we estimated the trehalose and residual water contents in the microenvironment of spin-dried cells. Individual cells with or without intracellular trehalose were embedded in a solid thin layer of extracellular trehalose after spin-drying. We found strong evidence suggesting that the residual water was bound at a 2:1 water/trehalose molar ratio in both the extracellular and intracellular milieus. Other than the water associated with trehalose, we did not find any more residual water in the spin-dried sample, intra- or extracellularly. The extracellular trehalose film exhibited characteristics of an amorphous state with a glass transition temperature of ∼22°C. The intracellular milieu also dried to levels suitable for glass formation at room temperature. These findings demonstrate a method for quantification of water and trehalose in desiccated specimens using confocal Raman microspectroscopy. This approach has broad use in desiccation studies to carefully investigate the relationship of water and trehalose content and distribution with the tolerance to drying in mammalian cells.  相似文献   

7.
The glassy state in corn embryos   总被引:3,自引:2,他引:1       下载免费PDF全文
The possibility is examined whether seeds may survive the desiccated state in part by vitrification, or the formation of a glassy state. Embryos excised from viable corn (Zea mays L.) seeds at low moisture contents show a series of low temperature first- and second-order phase transitions in the differential scanning calorimeter. These embryos produce normal seedlings if moistened. The thermal events can be duplicated almost entirely in both extracted lipids and purified commercial corn oil. They are therefore associated primarily with these bulk lipids, since membrane phospholipids are present in too small an amount to produce a detectable signal. When the bulk lipids have been extracted, a glass transition appears in the remaining material. At low water contents, it occurs above +40°C and systematically falls to below −60°C as the water content of the embryo rises to 20%. These data are consistent with our hypothesis that the desiccated state in seeds is a glassy state, and that imbibition of water reduces the glass transition temperature below ambient, allowing biochemical activity to resume.  相似文献   

8.
α-Amylase from mung beans (Vigna radiata) was immobilized on two different matrices, Amberlite MB 150 and chitosan beads. Maximum immobilization obtained was 72% and 69% in case of Amberlite and chitosan beads, respectively. The pH optima of soluble α-amylase were 5.6, whereas that for immobilized amylase on chitosan and Amberlite was 7.0. Soluble amylase and Amberlite immobilized amylase showed maximum activity at 65 °C, whereas chitosan immobilized amylase showed maximum activity at 75 °C. α-Amylase immobilized on Amberlite showed apparent Km of 2.77 mg/ml, whereas α-amylase immobilized on chitosan showed an apparent Km of 5 mg/ml. The Amberlite-amylase and chitosan-amylase showed a residual activity of 43% and 27%, respectively, after 10 uses. The loss of activity for free amylase after 100 days of storage at 4 °C was 70%, whereas that for Amberlite- and chitosan-amylases, under the same experimental conditions, the losses were 45% and 55%, respectively. The easy availability of mung bean α-amylase, the ease of its immobilization on low-cost matrices and good stability upon immobilization in the present study makes it a suitable product for further use in industrial applications.  相似文献   

9.
The objective of this study was to examine the composition and molecular structure of the endosperm cell walls (CW) derived from barley grain grown in three environments in Canada, and differing in grain hardness, protein, and total β-glucan contents. The endosperm CW were isolated from barley, cv. Metcalfe, grown in Davidson, SK (Sample A), Hythe, AB (sample B), and Hamiota, MB (sample C). The CW were sequentially extracted with water at 65 oC, saturated Ba(OH)2, again with water at 25 oC, and 1 M NaOH, resulting in fractions designated WE65, BaE, Ba/WE, and NaE, respectively. The monosaccharide analysis indicated the presence of β-glucans, arabinoxylans, and small amounts of arabinogalactans, glucomannans, and xyloglucans. Cellulose was detected in the CW remnants. The CW of sample A, exhibiting a lower grain hardness than sample B, contained the lowest amount of β-glucans, but the highest amount of arabinoxylans and the mannose-containing polysaccharides. The CW of sample C, characterized by very high protein content in the grain, contained the highest amount of β-glucans and the lowest amount of other polysaccharides. Polysaccharides in the CW of sample B, exhibiting the highest grain hardness, were characterized by the highest weight average molecular weights (Mw). β-Glucans in the CW of Sample B showed the highest ratio of DP3/DP4 and the longest cellulosic fragments in the polymeric chains. Of the three barley samples, arabinoxylans in the endosperm CW of sample A exhibited the lowest degree of branching, the highest amount of unsubstituted Xyl residues, and the highest ratio of singly to doubly substituted Xylp. The highest water solubility of the CW of sample C was associated with the highest concentration of β-glucans, the lowest DP3/DP4 ratio, and the lowest Mw of the polymeric constituents. Arabinoxylans with the lowest amount of doubly substituted but the highest amount of unsubstituted xylose residues and long sequences of unsubstituted xylan regions were found in the NaE fractions. The NaE fractions showed a high ratio of →4)-Glcp-(1→ to →3)-Glcp-(1→ linkages and some →4)-Manp-(1→ linkages, indicating a high level of long cellulosic regions in β-glucan chains and the presence of glucomannans.  相似文献   

10.
The objectives of this study were to investigate the moisture-induced protein aggregation of whey protein powders and to elucidate the relationship of protein stability with respect to water content and glass transition. Three whey protein powder types were studied: whey protein isolate (WPI), whey protein hydrolysates (WPH), and beta-lactoglobulin (BLG). The water sorption isotherms were determined at 23 and 45°C, and they fit the Guggenheim–Andersson–DeBoer (GAB) model well. Glass transition was determined by differential scanning calorimeter (DSC). The heat capacity changes of WPI and BLG during glass transition were small (0.1 to 0.2 Jg−1 °C−1), and the glass transition temperature (T g) could not be detected for all samples. An increase in water content in the range of 7 to 16% caused a decrease in T g from 119 down to 75°C for WPI, and a decrease from 93 to 47°C for WPH. Protein aggregation after 2 weeks’ storage was measured by the increase in insoluble aggregates and change in soluble protein fractions. For WPI and BLG, no protein aggregation was observed over the range of 0 to 85% RH, whereas for WPH, ∼50% of proteins became insoluble after storage at 23°C and 85% RH or at 45°C and ≥73% RH, caused mainly by the formation of intermolecular disulfide bonds. This suggests that, at increased water content, a decrease in the T g of whey protein powders results in a dramatic increase in the mobility of protein molecules, leading to protein aggregation in short-term storage.  相似文献   

11.
Low-moisture biopolymer-based systems are commonly encountered in food. Obviously, understanding the physical basis of their quality [texture, or performance over time or as a function of their composition (water or other added solutes)] is of primary importance. A polymer science approach using physical chemistry concepts based on physical state, phase transitions and molecular mobility can be applied to investigate the performances of food in particular versus moisture. Based on the example of starch-based samples and their texture property changes versus composition, the role of water and sucrose is considered through different aspects. The relations existing between the observed changes and physical state are investigated. While the motions associated with the glass transition were observed at high temperature, secondary relaxations are observed below Tg (at T β): T β decreased with water content and increased with increasing sucrose content. These local motions are suggested to contribute to the observed texture modifications versus water. Moreover, the stability of the glassy state was investigated by differential scanning calorimetry through the study of enthalpy relaxation (physical ageing). The amplitude of enthalpy relaxation decreased with both increasing sucrose and water content. All in all, this study strengthened the hypotheses that sub-Tg mobility could contribute to texture instability versus moisture or sugar content.  相似文献   

12.
Three synthetically produced glycolipids, N-(β-D-glucopyranosyl)-N-octadecyl-stearoylamide (OSGA), N-(β-D-glucopyranosyl-N-octadecyl-oleoylamide (OOGA), N-(β-D-galactopyranosyl)-N-octadecyl-lauroylamide (OLGA) have been studied in different mixtures with water by x-ray diffraction and dielectric measurements with microwaves at 9.4 GHz. The measurements were performed in the temperature range -50-70°C. X-Ray diffraction revealed a direct Lβ' → H transition at 20°C, 60°C, and 45°C depending on the glycolipid species but nearly not on the water content. The hexagonal phases are saturated at a water content of ≈20 wt%. The lamellar phase absorbs even less water (< 10 wt%). The dielectric data show that in the H phase the binding of water is stronger than in the Lβ' phase. In the temperature range below 0°C, OSGA and OOGA show a “subzero transition” due to the freeze-out of water in a separate ice phase. This transition can be seen in an abrupt decrease of the dielectric function because the dielectric response of ice is much smaller at microwave frequencies. OLGA does not show the subzero transition but an additional transition, hexagonal → distorted hexagonal at 60°C.  相似文献   

13.
The phase transitions in fully hydrated dipalmitoylphosphatidylcholine (DPPC) and DPPC/water/ethanol phases have been studied by lowangle time-resolved x-ray diffraction under conditions similar to those employed in calorimetry (scan rates 0.05-0.5°C/min and uniform temperature throughout the samples). This approach provides more adequate characterization of the equilibrium transition pathways and allows for close correlations between structural and thermodynamic data. No coexistence of the rippled gel (Pβ') and liquid-crystalline (Lα) phases was found in the main transition of DPPC; rather, a loss of correlation in the lamellar structure, observed as broadening of the lamellar reflections, takes place in a narrow temperature range of ~100 mK at the transition midpoint. Formation of a long-living metastable phase, denoted by Pβ'(mst), differing from the initial Pβ' was observed in cooling direction by both x-ray diffraction and calorimetry. No direct conversion of Pβ'(mst) into Pβ' occurs for over 24 h but only by way of the phase sequence Pβ'(mst) → Lβ' → Pβ'. According to differential scanning calorimetry (DSC), the enthalpy of the Pβ'(mst)-Lα transition is by ~5% lower than that of the Pβ'-Lα transition. The effects of ethanol (Rowe, E. S. 1983. Biochemistry. 22:3299-3305; Simon, S. A., and T. J. McIntosh. 1984. Biochim. Biophys. Acta 773:169-172) on the mechanism and reversibility of the DPPC main transition were clearly visualized. At ethanol concentrations inducing formation of interdigitated gel phase, the main transition proceeds through a coexistence of the initial and final phases over a finite temperature range. During the subtransition in DPPC recorded at scan rate 0.3°C/min, a smooth monotonic increase of the lamellar spacing from its subgel (Lc) to its gel (Lβ') phase value takes place. The width of the lamellar reflections remains unchanged during this transformation. This provides grounds to propose a “sequential” relaxation mechanism for the subgel-gel transition which is not accompanied by growth of domains of the final phase within the initial one.  相似文献   

14.
The kinetics and mechanisms of depolymerization of aqueous chitosan and alginate solutions at elevated temperatures have been investigated. Chitosan salts of different degree of acetylation (FA), type of counterions (-glutamate, -chloride) and degree of purity were studied. One commercially available highly purified sodium alginate sample with high content of guluronic acid (G) was also studied. Furthermore, the influence of oxygen, H+ and OH ions on the initial depolymerization rates was investigated. Depolymerization kinetics was followed by measuring the time courses of the apparent viscosity and the intrinsic viscosity. The initial rate constants for depolymerization were determined from the intrinsic viscosity data converted to a quantity proportional to the fraction of bonds broken. The activation energies of the chitosan chloride and chitosan glutamate solutions with pH close to 5 and the same degree of acetylation, FA = 0.14, were determined from the initial rate constants to be 76 ± 13 kJ/mol and 80 ± 11 kJ/mol, respectively.The results reported herein suggest that the stability of aqueous chitosan and alginate solutions at pH values 5–8 will be influenced by oxidative–reductive depolymerization (ORD) as the primary mechanism as long as transition metal ions are presented in the samples. Acid – and alkaline depolymerization will be the primary mechanisms for highly purified samples.  相似文献   

15.
Photoautotrophic cultivation of Euglena gracilis results in cells with high α-tocopherol content but the final cell concentration is usually very low due to the difficulty of supplying light efficiently to the photobioreactor. On the other hand, Euglena grows heterotrophically to high cell concentrations, using various organic carbon sources, but the α-tocopherol contents of heterotrophically grown cells are usually very low. Sequential heterotrophic/photoautotrophic cultivation, by which cells are grown heterotrophically to high cell concentrations and then transferred to photoautotrophic culture for accumulation of α-tocopherol was therefore investigated for efficient α-tocopherol production. In batch culture, using glucose as the organic carbon source, the cellular α-tocopherol content increased from 120 μg g−1 at the end of heterotrophic phase to more than 400 μg g−1 at the end of the photoautotrophic phase. By using ethanol as the organic carbon source during the heterotrophic phase, adding corn steep liquor as a nitrogen source and optimizing light supply during the photoautotrophic phase, the α-tocopherol content of the cells at the end of the photoautotrophic phase increased to 1700 μg g−1. A system consisting of a mini-jar fermentor (for the heterotrophic phase) and an internally illuminated photobioreactor (for the photoautotrophic phase) was then constructed for continuous sequential heterotrophic/photoautotrophic cultivation. The cells were continuously cultivated heterotrophically in the mini-jar fermentor and the effluent was continuously passed through the photobioreactor for α-tocopherol accumulation. In this way, it was possible to produce 7 g l−1 cells containing about 1100 μg α-tocopherol per g-cell continuously for more than 420 h. The continuous process resulted in α-tocopherol productivity of 100 μg l−1 h−1 which is about 9.5 and 4.6 times higher than those obtained in batch photoautotrophic culture and batch heterotrophic cultures, respectively.  相似文献   

16.
The crystal structure of the acyl complex of porcine pancreatic elastase with its peptidyl ester substrate N-acetyl-ala-ala-ala-methyl ester (Ac(Ala)3OMe) has been determined at 2.5 Å resolution. The complex was stabilized by exploiting the “glass transition” in protein dynamics that occurs at around −53 °C (220 K). Substrate was flowed into the crystal in a cryoprotective solvent above this temperature, and then the crystal was rapidly cooled to a temperature below the transition to trap the species that formed. The use of a flow cell makes the experiment a kinetic one and means that the species prior to the rate determining transition state has a chance to accumulate. The resulting crystal structure shows an acyl-enzyme intermediate in which the leaving group is absent and the carbonyl carbon of the C-terminal alanine residue is covalently bound to the gamma oxygen of the active site serine. The ester carbonyl shows no significant distortion from planarity, with the carbonyl oxygen forming one hydrogen bond with the oxyanion hole. The tripeptide is bound in an extended antiparallel β-sheet with main chain residues of the enzyme. The geometry and interactions of this acyl-enzyme suggest that it represents a productive intermediate. To test this hypothesis, the same crystal was then warmed above the glass transition temperature and a second data set was collected. The resulting electron density map shows no sign of the substrate, indicating hydrolysis of the intermediate followed by product release. This experiment provides direct evidence for the importance of dynamic properties in catalysis and also provides a blueprint for the stabilization of other short-lived species for direct crystallographic observation.  相似文献   

17.
Summary A wild coculture of obligately thermophilic bacteria, including only a single cellulolytic species Clostridium, ferments 2% crystalline cellulose and produces 4.6–5.1 g·l–1 of ethanol at 55°–60° C; that is, 0.96–1.1 moles of ethanol from 1 mole of glucose equivalent of cellulose degraded. However, the ethanol yield decreases with increasing cellulose concentration. Ethanolacetic acid ratio varies around 1 and cannot be influenced by substrate concentration. However, this ratio can be influenced by changing pH and temperature. For the ethanol production from cellulose, neutral and weekly alkaline media with a pH of 7.0–8.0 and a temperature of 55° C are optimal. Experiments in which the coculture was subjected to high ethanol concentrations showed that higher concentrations of added ethanol (up to 20 g·l–1) suppress cellulose degradation by 50% and inhibit the actual production of ethanol.  相似文献   

18.
Molecular simulation is becoming an important tool for both understanding polymeric structures and predicting their physical and mechanical properties. In this study, temperature ramped molecular dynamics simulations are used to predict two physical properties (i.e., glass transition temperature and thermal degradation temperature) of a previously synthesised and published telechelic benzoxazine. Plots of simulated density versus temperature show decreases in density within the same temperature range as experimental values for the thermal degradation. The predicted value for the thermal degradation temperature for the cured polybenzoxazine based on the telechelic polyetherketone (PEK) monomer was ca. 400°C, in line with the experimental thermal degradation temperature range of 450°C to 500°C. Mechanical Properties of both the unmodified PEK and the telechelic benzoxazines are simulated and compared to experimental values (where available). The introduction of the benoxazine moieties are predicted to increase the elastic moduli in line with the increase of crosslinking in the system.  相似文献   

19.
A low-molecular-weight water-soluble chitosan was obtained from high-molecular-weight crab chitosan using the enzyme preparation Celloviridine G20x. Optimum conditions for enzymatic hydrolysis were designed. The reaction should be performed for 4 h in a sodium-acetate buffer (pH 5.2) at 55°C and an enzyme to substrate ratio of 1 : 400. Fractional extraction of chitosan hydrolysate by aqueous ethanol (ethanol:distilled water) yielded fractions with molecular weights in the range 3.2–26.4 kDa.  相似文献   

20.
We have reported here on the structural polymorphism of lipid A, the “endotoxic principle” of bacterial lipopolysaccharide. For lipid A of rough mutant lipopolysaccharide from Salmonella minnesota and Escherichia coli, the three-dimensional supramolecular structures were determined with x-ray diffraction utilizing synchrotron radiation. The investigations were performed in the water concentration range 10 to 95% by weight, at [lipid A]:[Mg2+] molar ratios from 1:0 to 0.1:1, and in the temperature range from 20 to 70°C. These data were correlated with measurements of the β→α phase behaviour which was monitored with differential scanning calorimetry and Fourier-transform infrared spectroscopy. We found that the transition temperature of the acyl chains ranges—in the absence of Mg2+—from 45°C at high to 56°C at low water content, and—at an equimolar content of Mg2+—from 52°C at high to 59°C at low water concentrations. In the gel phase—in which the lipid A acyl chains are more disordered than those from saturated phospholipids—cubic phases are adopted at high water content (>60%) and at high [lipid A):[Mg2+] molar ratios. At low water contents, lamellar states are assumed exclusively. In the liquid crystalline state of lipid A, the hexagonal HII, state is adopted under all conditions. The structural variability of lipid A is highest at high water concentrations, and structural changes may be induced by only slight changes in temperature, water content, and Mg2+ concentration. Under physiological conditions, however, the lipid A assemblies exhibit a strong preference to cubic structures.  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