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.
Effect of six organic solvents—methanol, ethanol, propanol, dimethyl sulphoxide (DMSO), N,N-dimethyl formamide (DMF), and glycerol on the conformation and interaction of catalase and anticatalase antibodies were studied with the aim of identifying the solvents in which antigen–antibody interactions are strong. The antigen binding activity of the antibodies in the various organic solvents increased in the following order: ethanol < methanol < no organic solvent < propanol < DMSO < DMF < glycerol. The structure of both the antibody and the antigen molecule was affected significantly in 40% concentration of the organic solvents used in this study. Catalase activity was inhibited in DMSO. However, the enzyme was activated in DMF upto about 50% of its concentration.  相似文献   

2.
Using small-angle X-ray scattering, the behaviour of corn and potato starches during gelatinization, swelling, and rétrogradation was investigated. The scattering patterns were analysed on the basis of the fractal concept. The main scattering source from low moisture starches could be interpreted as a ‘surface fractal’ obeying a power law with an exponent of ca —4. When the starch swelled, the surface fractal characteristic was recognised only at very low angles, and on heating it disappeared at ca 80 °C. For gelatinised starches, the whole scattering pattern obeyed the power law with the power around -2.0. This result suggests that the physical arrangement of gelatinised starch molecules is a ‘mass fractal’, i.e. a self-similar structure, in nature. Further, it was found that the scattering of the retrograded starch showed a shoulder like peak superimposed the background scattering representing the mass fractal. The corresponding Bragg spacing was estimated to be 31˜15nm.  相似文献   

3.
The effect of milling parameters on the hydrolysis of starch during the mashing process was investigated. Hammer milling was compared against roll milling. Roll gap settings, roll speed, speed differential were also analysed, as well as comparing four- to six-roll milling. The parameter of differential speed was also studied through grist particle size distribution. Employing a 65 °C infusion type mashing process for the wort, the glucose and maltose concentrations of malts milled in different ways were analysed. Results showed that the glucose concentration in the wort after 45 min of mashing, obtained using a hammer mill, was the same as that achieved from roll milling in 60 min. For roller mill gap settings the 0.8 mm gap grist required 60 min of mashing to reach a glucose concentration of 3.46 g l−1, whereas the 0.1 mm gap grist achieved the same level of starch hydrolysis in almost half the time, around 30–35 min of mashing. The results regarding roll speed showed that the 300 and 700 rpm mashes required roughly 50 and 40 min, respectively. Comparable sugar concentrations in the 50 rpm mash were obtained in 60 min. Finally, the comparison between simulated four- and six-roll milling showed the latter yielded higher glucose concentrations.  相似文献   

4.
Gelatinization mechanism of potato starch   总被引:5,自引:0,他引:5  
The non-Newtonian behavior and dynamic viscoelasticity of potato starch (Jaga kids red ’90, 21.0% amylose content) solutions after storage at 25 and 4°C for 24 h were measured with a rheogoniometer. The flow curves, at 25°C, of potato starch showed plastic behavior >1.0% (w/v) after heating at 100°C for 30 min. A gelatinization of potato starch occurred above 1.0% at room temperature. A very large dynamic viscoelasticity was observed when potato starch solution (3.0%) was stored at 4°C for 24 h and stayed at a constant value with increasing temperature. A small dynamic modulus of potato starch was observed upon addition of urea (4.0 M) at low temperature (0°C) even after storage at 25 and 4°C for 24 h. A small dynamic modulus was also observed in 0.05 M NaOH solution. Possible models of gelatinization and retrogradation mechanism of potato starch were proposed.  相似文献   

5.
Various deoxycholic acid (DOCA)-modified-carboxymethylated (CM)-curdlan (DCMC) were synthesized and characterized by FTIR, 1H NMR and XRD. The degree of DOCA substitution (DS), as spectrophotometrically determined, was 2.1, 3.2, 4.1, or 6.3 DOCA groups per hundred sugar residues of CM-curdlan. The physicochemical properties of the self-assembled nanoparticals in aqueous media were investigated using 1H NMR, dynamic light scattering, zeta potential, transmission electron microscopy (TEM) and fluorescence spectroscopy. DCMC conjugates provided monodispersed self-assembled nanoparticles in water, with mean diameter decreasing from 192 to 347 nm with DOCA DS increasing. Moreover, the mean diameter also increased with decreasing pH in PBS. Zeta potential of DCMC self-assembled nanoparticles exhibited near −60 mV in distilled water and −26 to −36 mV in PBS, indicating these nanoparticles were covered with negatively charged CM-curdlan shells. The critical aggregation concentration (cac) of the DCMC were dependent on the degree of substitution (DS) of DOCA and were slightly lower in PBS than in distilled water. The TEM images demonstrated that these self-assembled nanoparticles were of spherical shape.  相似文献   

6.
Differential scanning calorimetry (DSC), rheological measurements and granule size analyses were performed to characterize the influence of phosphorylation substitution levels on the properties of cross-linked potato starch. Phosphorus oxychloride (POCl3) was used to produce the cross-linked potato starch. The levels of the reagent used for the reaction ranged between 40 and 5000 ppm (dwb). Storage (G′) and loss (G″) moduli were measured for a 5% (w/w) gelatinized starch dispersion stored at 20 °C for 24 h after heating at 85 °C for 30 min. The samples from 80 to 500 ppm were recognized as ‘strong gel'systems, whereas native potato starch showed ‘weak gel'behavior. Steady shear and dynamic viscoelastic properties of gelatinized starch dispersion were compared. Furthermore, granule mean diameter was measured by laser scattering for a 1% (w/w) dispersion heated at 85 °C for 30 min. The granules in the 100 ppm sample swelled to a maximum of about 2.6 times the native starch granule mean diameter.  相似文献   

7.
Raw cassava starch, having 74.94 and 0.44 g/100 g resistant starch type II and III (RS II and RS III), respectively, was autoclaved at 121 °C in water, 1, 10 or 100 mmol/L lactic acid. The formation of RS III was evaluated in relation to variable incubation temperature (−20 to 100 °C), incubation time (6–48 h) and autoclaving time (15–90 min). Negligible to low quantities of RS III (0.59–2.42 g/100 g) were formed from autoclaved starch suspended in 100 mmol/L lactic acid, whereas intermediate to high quantities (2.68–9.97 g/100 g) were formed from autoclaved starch suspended in water, 1 or 10 mmol/L lactic acid, except for treatments with water or 10 mmol/L lactic acid incubated at 100 °C for 6 h (1.74 g/100 g). Autoclaving times corresponding to maximum RS III contents were 15 and 45 min for water and 10 mmol/L lactic acid, respectively. Whereas, the RS III fractions from cassava starch suspended in water had melt transitions between 158 and 175 °C with low endothermic enthalpies (0.2–1.6 J/g), the thermal transitions of the acid treated samples were indistinct.  相似文献   

8.
Microencapsulations of cumin oleoresin by spray drying using gum arabic, maltodextrin, and modified starch (HiCap® 100) and their ternary blends as wall materials were studied for its encapsulation efficiency and stability under storage. The microcapsules were evaluated for the content and stability of volatiles, and total cuminaldehyde, γ-terpinene and p-cymene content for six weeks. Gum arabic offered greater protection than maltodextrin and modified starch, in general, although the order of protection offered was volatiles > cuminaldehyde > p-cymene > γ-terpinene. A 4/6:1/6:1/6 blend of gum arabic/maltodextrin/modified starch offered a protection, better than gum arabic as seen from the t1/2, i.e. time required for a constituent to reduce to 50% of its initial value. However protective effect of ternary blend was not similar for the all the constituents, and followed an order of volatiles > p-cymene > cuminaldehyde > γ-terpinene.  相似文献   

9.
Cross-linked waxy maize (CWM) starch dispersions (STDs) of concentration 50 g kg−1 were heated in sucrose solutions containing 0–600 g kg−1 (g sucrose/kg dispersion) at 85 °C at low shear and in intermittently agitated cans at 110 °C. The STDs heated in 0–300 g kg−1 sucrose exhibited antithixotropic behavior, while those heated in 400–600 g kg−1 sucrose exhibited thixotropic behavior. The mean starch granule diameter of the starch dispersions did not show strong dependence on sucrose concentration. The dispersions, especially those with high sucrose concentrations and heated at 110 °C, exhibited G′ versus frequency (ω) profiles of gels. The STDs exhibited first normal stress differences that increased in magnitude with the concentration of sucrose. Values of the first normal stress coefficient of canned dispersions calculated from dynamic rheological data plotted against ω and experimental values plotted against shear rate of some of the STDs overlapped.  相似文献   

10.
Binary systems of polymers often display spherulitic morphologies after cooling from the melt, but these phenomena have rarely been reported among food polymers of native-size. Here we report the observation of spherulitic and other morphologies in gelatinized maize starch. The morphology could be manipulated by choosing polymer compositions and kinetic regimes. Spherulites (10 μm diameter) formed from gelatinized high-amylose maize starches and purified amylose at cooling rates of order of magnitude 100 °C/min. They were more numerous and exhibited a higher melting point the greater the ratio of amylose to amylopectin. Rapid cooling rates (150–500 °C/min) resulted in a more even distribution of smaller spherulites. Very rapid (liquid nitrogen quench) or slow (0.1–1 °C/min) cooling rates resulted in mixed morphology, as did addition of 15 or 60% (w/w) sucrose to a 10% (w/w) dispersion of high-amylose starch (HAS). Spherulites were observed in aqueous suspensions of high-amylose maize starch between 5 and 30% (w/w). Lower starch concentrations resulted in a broader size distribution and spherulites of more distinct shape. WAXS patterns of B-type were observed. Negatively birefringent spherulites predominated, but positive spherulites were found. The spherulite melting range overlapped with that for amylose–lipid complex. Evidence indicated that micro-phase separation takes place when a holding period at 95 °C follows gelatinization at 180 °C. Despite the high maximum temperature of treatment (180 °C) there was evidence for a memory effect in samples of 30% HAS. Spherulite morphology closely resembled that of native starch granules in very early stages of development.  相似文献   

11.
Compatibility, flow and visco-elastic properties of a pregelatinised maize starch mixed with κ-carrageenan were investigated. After cooking of the pregelatinised starch, some undissolved granules remained in solution. Aqueous mixtures of κ-carrageenan and starch were studied at 60 °C and 20 °C by combining rheological measurements and microscopic observations under conditions allowing gelation of carrageenan and non-gelation of starch. The viscometric study of mixed dilute solutions of amylose from pregelatinised starch and carrageenan showed that the components are slightly incompatible. Mixture viscosity and elastic modulus were studied at 60 °C in details as a function of mixture composition for a total polymer concentration of 3%; both were found to be significantly higher than the corresponding theoretical additive values. This finding was interpreted by starch granules excluded volume effect. At 20 °C, no noticeable increase of mixture elastic modulus was found as compared with the additive value. The absence of the synergistic effect is supposed to be due to the formation of highly inhomogeneous gels with agglomerates of undissolved granules.  相似文献   

12.
Acylation reactions of naringin with palmitic acid were performed by a lipase after formation of highly concentrated homogeneous solutions. Their initial naringin concentration was 840–950 mM, which is 20–60 times greater than that in organic solvent media. The overall productivity of highly concentrated solutions was more than 15 times greater than those of organic phase media. The addition of DMSO (20–40%, w/w) to substrate mixtures lowered the melting temperature of a naringin–palmitic acid mixture (1:1 molar ratio) to about 40 °C. Reactions at 80 °C apparently followed Michaelis–Menten kinetics despite extremely high substrate concentrations. As the temperature increased from 60 °C to 80 °C, the apparent viscosity of the highly concentrated solution decreased remarkably from 4.31 Pa s to 0.063 Pa s. An activation energy of 7.65 kcal/mol obtained in a range of 60–75 °C suggests a diffusion-control. On the other hand, an activation energy of 17.09 kcal/mol in a range of 75–90 °C indicates a reaction-control. The highest product conversion yield of 33% (mol/mol) was obtained in a 10 h reaction at 80 °C. Addition of activated molecular sieves to the highly concentrated solution increased the product conversion yield by 7% (mol/mol), suggesting that the original equilibrium was disrupted by removing water and then a new equilibrium was reached.  相似文献   

13.
Cationic starch nanoparticles were obtained by aqueous polyelectrolyte complex formation between cationic quaternary ammonium substituted starches and anionic sodium tripolyphosphate. The formation of nanosized starch particles of spherical shape was verified by dynamic light scattering and scanning electron microscopy measurements. The cationic starch nanoparticles of different constitution and containing various contents of free quaternary ammonium groups were produced and their zeta potential was modulated between +4 mV and +34 mV by varying polycation/polyanion ratio. Furthermore, the polyelectrolyte complex formation was confirmed by differential scanning calorimetry and FTIR analyses. The thermal stability of cationic starch nanoparticles increased with the introduction of polysalt into polyelectrolyte complex. The solubilization capacity of nanoparticles was varying with the concentration and composition as revealed by fluorescence probe experiments. The capability to accommodate hydrophobic pyrene quest molecule was decreasing with the increasing number of cationic groups in cationic starches and little depended on polyanion/polycation ratio in starch nanoparticles.  相似文献   

14.
A novel cellulose-based hydrophobically associating water-soluble polymer-cellulose octaonate sulfate (COS) was synthesized in this paper. The basic physico-chemical properties such as surface tension and the critical aggregation concentration (cac) were measured by the conventional Wilhelmy plate method. The obtained cac value was compared with environmental scanning electron microscopy (ESEM) and Rheology data. All these results indicated that cac of this amphiphilic polymer was between the range 0.04 and 0.2 wt% and the corresponding surface tension was around 55 mN/m. The conformation of aggregates and size of particles in aqueous solution were carefully investigated by ESEM and dynamic laser scattering (DLS) measurements. When the concentration is around 0.04 wt%, loose aggregates are formed; around 0.5 wt%, network structure formed. DLS results indicated that average size of particle was increased from 54.7 nm to 73 nm and finally to 168.1 nm with the increase of concentration from 0.04% to 0.1% and even to 0.2%. These results suggested that almost all of micelles in aqueous solution aggregated at the experimental concentration range 0.04–0.5 wt%. Rheological properties of this polymer were similar to hydrophobically associating polymers’ (HMP). As the shear rate increased, the solution passed through a shear-thickening region before exhibiting a sharp decrease in viscosity, eventually exhibited Newtonian behavior.  相似文献   

15.
Effects of proportions of neutral detergent fibre (aNDFom) and starch, as well as their degradation rates, on rumen fermentation were tested using an in vitro rumen simulation system (SIMCO). The in vitro system was designed to simulate selective particle retention and had an average fluid volume of 1150 ml with a liquid dilution rate of approximately 0.07 h−1. Two types of hay (aNDFom sources) and two types of starch were each included at two different levels in the diet and were examined in an experiment following a 2×2×2 factorial arrangement of treatments (eight diet combinations). The hay was either late-cut timothy (Phleum pretense L.) or early cut meadow grass (Poa pratensis L.), with ruminal in situ aNDFom digestion rates of 0.03–0.04 and 0.07–0.08 h−1, respectively. The two starch types were raw (R) and cooked (C) potato starch with previously determined in vitro ruminal digestion rates of 0.04 and 0.20 h−1, respectively. The starch levels were 300 and 600 g/kg diet dry matter (DM) with the remaining being hay (282–682 g/kg DM) and peptone (14–111 g/kg DM). The aNDFom level varied among the diets with different starch levels and hay types. The peptone acted as a source of peptides and, together with ammonia salts from buffer, was used to balance the N contents of the diets. The feeding level for each of the eight vessels was 28 g DM/d. Two 10-day simulations were made with the system. The average pH was higher (P<0.05) for all treatments with raw potato starch (6.19) versus cooked starch (6.07). Protozoa scores, on a qualitative scale, declined faster at the higher starch level. The aNDFom digestibility was, as expected, higher (P<0.001) for meadow hay (0.57) than timothy (0.32), and was also higher (P<0.001) at the lower starch level (0.54) versus the higher (0.35). Microbial protein production efficiency (mg microbial N/g organic matter truly digested) was higher for the faster degrading aNDFom (P<0.01) and starch (P<0.05) sources, but was not affected by starch level. Cooked starch resulted in a lower acetate proportion (449 mmol/mol versus 591 mmol/mol VFA; P<0.001) but higher proportions of propionate (297 mmol/mol versus 236 mmol/mol VFA; P<0.001), and butyrate (169 mmol/mol versus 127 mmol/mol VFA; P<0.01). Butyrate increased with starch level (127 mmol/mol versus 169 mmol/mol VFA; P<0.01), and was also higher for meadow hay versus timothy (168 mmol/mol versus 128 mmol/mol VFA; P<0.01). Interactions between the treatments demonstrate that the response in VFA pattern to starch level is dependent on starch and aNDFom sources. Substrates such as starch and aNDFom are fermented differently depending on their rates of ruminal degradation.  相似文献   

16.
Native new cocoyam starch (nNCS) was subjected to annealing (aNCS) and heat moisture treatment at 18% moisture level (h18NCS), 21% moisture level (h21NCS), 24% moisture level (h24NCS) and 27% moisture level (h27NCS) as hydrothermal treatments. Scanning electron and light microscopy revealed round and polygonal shapes with sizes ranging from 15 to 40 μm for native and modified starches. nNCS showed “A” pattern X-ray diffraction and no significant differences were observed in the X-ray pattern of the modified starches. Swelling power and solubility reduced following heat moisture treatment. At all pH studied (2–12), unmodified new cocoyam starch exhibited higher swelling capacity and solubility than the modified derivatives. Hydrothermal modifications improved water absorption capacity but reduced oil absorption capacity. Pasting temperature of native starch shifted to higher values following annealing and heat moisture treatment. Hot paste viscosity (Hv), viscosity after 30 min holding at 95 °C (Hv30) and cold paste viscosity (Cv) reduced after annealing and heat moisture treatment. The result also indicates that hydrothermal treatments reduced the tendency for setback. As the number of days of storage of starch paste increased from 1 to 10, light transmittance of all the starches reduced but marked reduction of light transmittance was observed in native starch. DSC studies revealed increase in gelatinization temperature following annealing and heat moisture treatment. Starch hydrothermal modifications reduced retrogradation as enthalpies of regelatinization reduced following modifications. The regelatinization peak in the second day scanning shifted to lower temperature than the gelatinization peak in first run heating DSC curve for all samples. The regelatinization peak also became larger and shifted to higher temperature range when the storage days increased from 2 to 7.  相似文献   

17.
The adsorption of bromelain from an aqueous solution by polyacrylic acid (PAA)-bound iron oxide magnetic nanoparticles was studied. The magnetic composite nanoparticles were shown to be efficient for the separation of bromelain. Except at pH <3, the adsorption of bromelain increased with the decrease in solution pH and reached almost 100% at pH 3–5. The adsorbed bromelain could be desorbed by the addition of KCl and complete desorption was achieved at pH 7 when [KCl]>0.6 M. The adsorption behaviour followed the Langmuir isotherm with a maximum adsorption amount of 0.476 mg/mg and a Langmuir adsorption equilibrium constant of 58.4 ml/mg at pH 4 and 0.1 M phosphate. In addition, it was notable that both the adsorption and desorption of bromelain were quite fast and could be completed in about 1 min due to the absence of internal diffusion resistance. Bromelain retained 87.4% activity after adsorption/desorption.  相似文献   

18.
Effects of Mg2+ ions on thermally induced conformational transitions in the synthetic poly(dA)·poly(dT) and poly(dA)·2poly(dT) were studied in the buffered solutions (pH 6.9), containing 0.1 or 1 M NaCl at polynucleotide concentration of 0.1–0.3 mM (in nucleic bases). The experiments consist of measurements of the UV absorption and intensity of conventional visible static light scattering. The diagram of conformational transitions in the poly(dA)–poly(dT)–Mg2+ system was constructed on a basis of experimental data obtained. Anomalously strong light scattering, like critical opalescence, has been revealed at 0.1 M NaCl and [Mg2+]≥20 mM in the melting range of both polynucleotides, which eventually disappeared after the completion of polymer strands separation. The effect presumably is caused by a fluctuation process of polymer strands complexing which arises at a certain concentration of Mg2+ ions.  相似文献   

19.
Three commercial chitosans with a degree of acetylation of 25–30% were studied by light scattering (static and dynamic), analytical ultracentrifugation (sedimentation velocity and sedimentation equilibrium), and capillary viscometry in 0.02 M acetate buffer/0.1 M NaCl, pH 4.5. The molecular masses obtained by sedimentation equilibrium measurements or sedimentation and diffusion coefficients according to the Svedberg equation agreed well or fairly well with those from static light scattering whereas the molecular masses calculated via the Scheraga–Mandelkern equation were found too low by almost 50%. The various Mark–Houwink type relationships suggested a nearly free-draining flexible worm-like chain. A prolate ellipsoid of revolution with an axial ratio a/b25 was shown to be a hydrodynamically equivalent body of the flexible worm-like chain that had been derived from static light scattering. The findings illustrate the fact that a hydrodynamically strongly asymmetric shape need not mean a strongly elongated shape of the molecules in reality.  相似文献   

20.
A novel citric acid (CA)–glycerol co-plasticized thermoplastic starch (CGTPS) was prepared by melt blending. The CA content varies from 10% to 40 wt%. Result from Fourier Transform Infrared spectroscopy (FTIR) show that partial esterification occurred during blending. The degrees of substitution and esterification increased as the CA content increased. Results from intrinsic viscosity measurement, laser light scattering (LLS), and FTIR demonstrate the molecular weight of starch decreased as the CA percentage increased. The weight average molecular weight (Mw) of CGTPS with 20 wt% CA was only one-tenth of that without CA under the same processing conditions. Crystal type and crystallinity changes as a function of CA were recorded by X-ray diffraction (XRD). Thermal stability and the glass transition temperature (Tg) were detected by thermogravimetric (TG) and differential scanning calorimeter (DSC). Compared to the traditional GTPS, the novel CGTPS exhibits the special characters of partial esterification, low molecular weight and stronger interaction between starch and plasticizers. These new properties can be expected to prevent retrogradation, promote compatibility with polyesters, improve the processing ability, and adjust the degradation properties.  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