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 46 毫秒
1.
The structure and interaction of κ-carrageenan and locust bean gum (LBG) has been studied using rheology, cryo-SEM, conductivity and syneresis characterization. The rheological behaviour of the binary system has been characterized using both compression and shear measurements. Elimination of slip in the shear measurements yields G′ values of the order 10,000–30,000 Pa for a 1% κ-carrageenan gel in 0–0.2 M added KCl. These values are higher than previously reported. No synergistic peak was found with the addition of LBG as has been previously reported. The measured modulii for these gels yields a Poisson's ratio of 0.5. Compression rupture stress and strain were also monitored. The rupture measurements do show a synergistic peak indicating that the interaction does occur and is important at high strain amplitudes. The gel points as determined by conductivity for these systems show a decrease in temperature with increasing LBG concentration, which is consistent with rheological measurements. Syneresis results are reported for the range of κ-carrageenan/LBG ratios. The syneresis shown by the mixtures is the same as that shown by the same concentration of κ-carrageenan. Structures of the gels as determined by cryo-SEM are also reported. Characteristic length scales in these systems are of the order of tens of microns and show little change with LBG concentration. The reduction in the characteristic length scale with increasing LBG concentration is discussed in terms of the rheological behaviour.  相似文献   

2.
The role of σB in the Staphylococcus aureus heat-shock induced thermotolerance was investigated. Survival curves at 58 °C of S. aureus strain Newman and its isogenic ΔsigB mutant were obtained for native and heat-shocked cells (45 °C for 5–120 min) in exponential and stationary phase of growth. The magnitude of the acquisition of thermotolerance at 58 °C depended on the growth phase and on the duration of the heat shock. Stationary growth phase cells were always more heat tolerant than exponentially growing cells and thermotolerance increased with heat-shock duration up to 120 min. S. aureus cells were able to increase their heat tolerance in the absence of the σB factor. In stationary phase, whereas in the parental strain the thermotolerance was increased by a factor of 12 after a heat shock of 120 min at 45 °C (δ values at 58 °C for native and heat-shocked cells were 0.63 and 7.22 min, respectively), in the mutant strain it increased 43 fold (δ values 0.09 and 3.87 min). The addition of chloramphenicol to the adaptation medium resulted in a lower increase in heat tolerance but did not prevent it completely, suggesting that S. aureus can partially increase its thermotolerance without “de novo” protein synthesis. Both the number of non-damaged cells and the proportion of cells able to repair sublethal damage were higher for heat-shocked cells.  相似文献   

3.
Mixed gels, formed by κ-carrageenan, and sodium caseinate were studied by differential scanning calorimetry (DSC) and rheometry. DSC showed that during gelation (i.e. cooling) the thermal behaviour of κ-carrageenan was almost uninfluenced by the presence of sodium caseinate. Thus the interaction of κ-carrageenan with sodium caseinate has little (or no) effect on the carrageenan's coil-to-helix transition. In contrast, during melting, added sodium caseinate strongly modified the thermal behaviour. The DSC peak became progressively broader with addition of sodium caseinate, indicating that the junction zones are highly heterogeneous in the mixed gel. Rheometry showed that sodium caseinate strongly influences the storage modulus (G′). In experiments in which the concentration of sodium caseinate was fixed and that of κ-carrageenan varied, plots of G′ vs. concentration of κ-carrageenan were biphasic, with an abrupt change in slope at a concentration that increased linearly with the concentration of sodium caseinate. When the concentration of κ-carrageenan was constant and that of sodium caseinate varied, G′ as a function of concentration of sodium caseinate passed through a minimum. This behaviour could be modelled quantitatively, by assuming that: (a) the sodium caseinate adsorbs κ-carrageenan, but with a limited adsorptive capacity; (b) sodium caseinate aggregates (sub-micelles) with adsorbed κ-carrageenan can associate via interaction between free ends of adsorbed κ-carrageenan chains and form a gel network; and (c) the contributions to G′ from the sodium caseinate–κ-carrageenan network and the network formed by κ-carrageenan alone are additive. At low κ-carrageenan to sodium caseinate ratios, the sodium caseinate and κ-carrageenan combine to form a mixed gel. As the ratio of κ-carrageenan to sodium caseinate increases, the sodium caseinate becomes saturated and no further association with κ-carrageenan can occur—the increase in G′, as further κ-carrageenan is added, comes from a gel network formed by κ-carrageenan alone.  相似文献   

4.
There is an interest in developing protein based thickening agents for nutritional considerations. A procedure to convert whey protein concentrates or isolates into a pH modified cold-thickening ingredient was developed. Concentration effects on thickening mechanism of this whey protein ingredient were studied with a β-lactoglobulin model system at the pH of the modification procedure, 3.35. In this study, concentration effects on thermal aggregation of β-lactoglobulin were studied at low pH using capillary and rotational viscometry, transmission electron microscopy (TEM), and high performance liquid chromatography coupled with multi-angle laser light scattering (HPLC-MALS). From the results of capillary viscometry, a critical concentration (Cc  6.9% w/w) was identified below which no significant thickening functionality could be achieved. Microscopy revealed formation of flexible fibrillar network at pH 3.35 during heating at all concentrations. These flexible fibrils had a diameter of about 5 nm and persistence length of about 35 nm as compared to more linear and stiff fibrils formed at pH 2 and low ionic strength conditions. Under similar heating conditions at concentration above Cc, larger aggregates similar to microgels were observed compared to the concentration below Cc, where isolated fibrils with an average contour length of about 130 nm were observed. These microgels and apparently stronger interactions between aggregates at concentrations above Cc were seemingly responsible for thickening functionality of heated β-lactoglobulin solutions and subsequently modified powders. Further investigation of β-lactoglobulin aggregation at this pH may provide capability to mechanistically tailor the functional attributes of modified ingredients.  相似文献   

5.
Equivalent sweetness of aspartame relative to two sucrose concentrations (10% and 20% w/w) were determined in water and in hydrocolloids gels. The influence of the texture of three hydrocolloids gelled systems—gellan gum, κ-carrageenan, and κ-carrageenan/locust bean gum (LBG)—at two gums concentrations (0.3% and 1.2% w/w) on the equivalent sweetness of aspartame were then studied. For the three gelled systems, the increase in hydrocolloid concentration produced a significant increase in the true rupture stress and in the deformability modulus values. For both κ-carrageenan and mixed gels the true rupture strain values increased when increasing hydrocolloid concentration while for gellan gels, decreased. For the same hydrocolloid concentrations the κ-carrageenan/LBG gels showed the largest strain at rupture and gellan gels the smallest (most brittle). For both soft (0.3% gum) and hard (1.2% gum) gellan gels and κ-carrageenan gels, the concentrations of aspartame needed to deliver a sweetness intensity equivalent to that of gels with 10% sucrose (0.079–0.087% w/w) were similar to those obtained for aqueous solutions (0.084% w/v). For hard κ-carrageenan/LBG gels the corresponding concentration of aspartame was slightly lower. For all gelled systems the concentrations of aspartame needed to deliver a sweetness intensity equivalent to that of gels with 20% sucrose were higher for soft gels than for hard gels.  相似文献   

6.
Oat β-glucan hydrolysates with different molecular weights were prepared and their physicochemical, hypocholesterolemic, and weight-reducing characteristics were evaluated. The enzymatic hydrolysis by cellulase caused a decrease in the molecular weight of oat β-glucan (1450–370 × 103 g/mol), which also affected swelling power and bile acid/fat binding capacities. In addition, mice were fed high-fat diet supplemented with β-glucans with three different molecular weights (1450, 730, 370 × 103 g/mol). The diets treated with β-glucans significantly reduced the body weight of the mice. However, the molecular weight of β-glucans did not appear to significantly affect the serum lipid profile.  相似文献   

7.
Using differential scanning calorimetry, the gel-sol transition temperatures were compared for KCl-precipitated κ-carrageenan and Philippines natural grade (PNG) κ-carrageenan. The latter contains cellulose, which is associated with the carrageenan in seaweed. This leads to lower gel-sol transition temperatures. More heat absorption (-ΔHm = 79 kJ/mol) is required for the formation of junction zones for the PNG κ-carrageenan than the KCl-precipitated κ-carrageenan (-ΔHm = 40 kJ/mol). This behaviour can be attributed to the cellulose interspersing between κ-carrageenan aggregates in the PNG κ-carrageenan gels.  相似文献   

8.
The purpose of this study was to determine the influence of neutral cosolvents on the formation and properties of biopolymer nanoparticles formed by thermal treatment of protein–polysaccharide electrostatic complexes. Biopolymer particles were formed by heating (85 °C, 20 min) an aqueous solution containing a globular protein (β-lactoglobulin) and an anionic polysaccharide (beet pectin) above the thermal denaturation temperature (Tm) of the protein under pH conditions where the biopolymers formed electrostatic complexes (pH 5). The impact of two neutral cosolvents (glycerol and sorbitol) on the self-association of β-lactoglobulin and on the formation of β-lactoglobulin–pectin complexes was examined as a function of solution pH (3–7) and temperature (30–95 °C). Glycerol had little impact on the pH-induced self-association or aggregation of the biopolymers, but it did increase the thermal aggregation temperature (Ta) of the protein–polysaccharide complexes, which was attributed to its ability to increase aqueous phase viscosity. Sorbitol decreased the pH where insoluble protein–polysaccharide complexes were formed, and greatly increased their Ta, which was attributed to its ability to increase Tm, alter biopolymer–biopolymer interactions, and increase aqueous phase viscosity. This study shows that neutral cosolvents can be used to modulate the properties of biopolymer nanoparticles prepared by thermal treatment of protein–polysaccharide electrostatic complexes.  相似文献   

9.
The susceptibility of an industrial α-lactalbumin concentrate to cross-linking with a microbial transglutaminase from Streptoverticillium mobaraense was investigated. At a protein concentration of 0.5% w v−1, the maximum cross-linking was observed at 50°C, pH 5 and at 5 h of incubation time. Results from sodium dodecyl sulphate (SDS)-polyacrylamide gel electrophoresis showed that most of the monomeric form of α-lactalbumin was converted to polymers too large to move into the gel matrix. Addition of ethylenediamine tetraacetic acid or SDS prior to the incubation of protein–enzyme mixture, further enhanced the transglutaminase reaction with the industrial α-lactalbumin. Results from reverse phase chromatography indicated that cross-linking caused a broadening of the α-lactalbumin peak with little change in the average hydrophobicity of the protein. In contrast to the reported results on pure α-lactalbumin, the industrial α-lactalbumin concentrate showed considerable cross-linking with transglutaminase even without the reduction of the disulphide bonds. This difference was attributed to the partially unfolded secondary structures in the industrial α-lactalbumin concentrate.  相似文献   

10.
The effects of molecular size and fine structure of mixed-linkage cereal (1→3), (1→4) β- -glucans (β-glucans) on their cryogelation behavior were investigated. Values of apparent molecular weight (Mw) for oat β-glucans ranged between 65 and 200×103, whereas the respective values for both barley and wheat β-glucan preparations were about 200×103. The fine structure of cereal β-glucans, as assessed by high-performance anion-exchange chromatography of the cellulosic oligomers released by the action of lichenase, revealed differences in the relative amounts of 3-O-β-cellobiosyl- -glucose (DP3) and 3-O-β-cellotriosyl- -glucose (DP4) units only among the different genera of cereals; the weight percent of DP3 units estimated as 67.1, 63.3, and 55.3–55.8% for wheat, barley, and oat β-glucans, respectively. Aqueous β-glucan solutions (1–3% w/v) were subjected to 12 freezing (−18 °C for 24 h) and thawing (5 °C for 24 h) cycles. The phenomenological appearance of the gelled materials obtained after this process as well as the yield of cryostructurates were influenced by the initial solution concentration, the number of freeze–thaw cycles, as well as by the molecular features of the β-glucans. Such effects were unraveled by studying the cryogelation process with differential scanning calorimetry (DSC), small strain dynamic rheometry, and large deformation mechanical measurements. For the cereal β-glucan cryogels the storage modulus, G′, increased and the tan δ decreased with decreasing polysaccharide molecular size and with increasing initial solution concentration, number of freeze–thaw cycles, and trisaccharide segments in the polymeric chains. The apparent melting enthalpy values (ΔH) of β-glucan cryostructurates, as determined from the DSC endothermic peaks, increased with decreasing molecular size and with increasing amount of cellotriose units, but they were independent of the number of freeze–thaw cycles. The DSC melting temperature of the gel network was found to increase with the molecular size and amount of DP3 units of β-glucans. Moreover, large deformation mechanical tests (compression mode) revealed an increase in strength of cereal β-glucan cryogels with increasing molecular size and decreasing trisaccharide units in the polysaccharide preparation.  相似文献   

11.
Enzymatic modification of proteins, in order to produce functional materials such as hydrogels, adhesives and films via cross-linked networks or scaffolds of proteins, is a constantly evolving technology to create tailored micro- and nanostructured materials for food, cosmetic, and medical applications. For the successful utilization of oxidoreductases or transferases such as tyrosinases and transglutaminases, respectively, it is crucial to understand the action of these enzymes on protein substrates. In this study, cross-linking of the milk protein β-casein by Trichoderma reesei tyrosinase (TrTyr) was studied using size-exclusion chromatography (SEC) equipped with multi-angle light scattering (MALLS) and ultraviolet/visible (UV/Vis) detectors in order to determine the molecular mass (MM), radius of gyration (RG) and degree of polymerization (DP) of the reaction products. Sodium dodecyl sulphate-polyacrylamide gel electrophoresis (SDS-PAGE) was used to detect early polymerization states. The widely used Streptoverticillium mobaraense transglutaminase (Tgase) was used for comparison to tyrosinase from T. reesei. The results showed that cross-linking of β-casein by these two different types of enzymes resulted in the formation of polymerized reaction products with MM ranging from 500 to 1700 kg mol−1 depending on the enzyme dosage and incubation time. The DP varied from 21 to 71, respectively. In the case of TrTyr the polymerized reaction products were slightly colored, and formation of the covalent cross-linking of β-casein could be monitored by UV/Vis as a function of incubation time.  相似文献   

12.
Ivano De Noni   《Food chemistry》2008,110(4):897-903
The release of β-casomorphin-5 (BCM5) and β-casomorphin-7 (BCM7) was investigated during simulated gastro-intestinal digestion (SGID) of bovine β-casein variants (n = 3), commercial milk-based infant formulas (n = 6) and experimental infant formulas (n = 3). SGID included pepsin digestion at pH 2.0, 3.0 and 4.0 and further hydrolysis with Corolase PP™. β-Casein (β-CN) variants were extracted from raw milks coming from cows of Holstein-Friesian and Jersey breeds. Genomic DNA was isolated from milk and the β-CN genotype was determined by a PCR-based method. Phenotype at protein level was determined by capillary zone electrophoresis in order to ascertain the level of gene expression. Recognition and quantification of BCMs involved HPLC coupled to tandem MS. Regardless of the pH, BCM7 generated from variants A1 and B of β-CN (5–176 mmol/mol casein) the highest amount being released during SGID of form B. As expected, the peptide was not released from variant A2 at any steps of SGID. BCM5 was not formed in hydrolysates irrespective of either the genetic variant or the pH value during SGID. Variants A1, A2 and B of β-CN were present in all the commercial infant formulae (IFs) submitted to SGID. Accordingly, 16–297 nmol BCM7 were released from 800 ml IF, i.e. the daily recommended intake for infant. Industrial indirect-UHT treatments (156 °C × 6–9 s) did not modify release of BCM7 and, during SGID, comparable peptide amounts formed in raw formulation and final heat-treated IFs.  相似文献   

13.
Ovalbumin is a key ingredient in many food products due to its ability to form a three-dimensional network after heat denaturation, that is, to form gels. We have explored two ways of modifying gels based on ovalbumin: (1) the addition of carrageenans, and (2) manothermosonication (MTS) treatments of ovalbumin suspensions prior to gelation. Ovalbumin gels showed elastic modulus (G′) values of 1216 Pa after a heating and a cooling period. The addition of kappa, lambda, and iota carrageenan enhanced G′ values by 6, 1.7, and 1.7 fold, respectively. Kappa carrageenan was not only the most effective carrageenan in the improvement of ovalbumin gelation but changed also the gelation process at the end of the cooling phase. MTS treatments reduced G′ values of ovalbumin gels by about 50% measured immediately after MTS treatments but only by ca. 25% after 24 h, which reveals that ovalbumin modifications introduced by ultrasound, whatever they are, are reversible. This reduction does not preclude the use of manothermosonicated liquid egg as gelling agent.  相似文献   

14.
The literature reports an optimum NaOH concentration for the alkaline cleaning of whey deposits or gels; at NaOH concentrations higher than this optimum, cleaning proceeds much more slowly. Although this phenomenon is of great importance in the cleaning of dairy equipment, no conclusive physical explanation has yet been presented. In this study, we present strong evidence that the dissolution rate is affected by the equilibrium-swelling ratio in β-lactoglobulin (βLg) gels. The swelling ratio is greatly reduced in the presence of salts due to the polyelectrolyte screening effect of the cations. This has been observed in free-swelling βLg gels using gravimetrical analysis and in the uniaxial swelling of WPC gel deposits using fluid dynamic gauging. At high dissolution pH (>13.3), the high Na+ concentration reduces swelling in spite of the high surface charge of the protein. It is proposed that the reduction of the free volume inside the gel impedes the transport of the protein aggregates out of the NaOH penetration zone. We have also observed that the final dissolution rate of gels pre-soaked in 1 M NaOH or NaCl is similar, despite the difference in pH, and much lower than for untreated gels: the high Na+ concentration in the soaked gels hinders swelling, inhibiting the disentanglement of the protein clusters regardless of the high pH.  相似文献   

15.
Purified (1→3)(1→4)-β- -glucans (β-glucans) from oat and barley with broad molecular weight (MW) distribution were separated into seven fractions using gradient precipitation with ammonium sulfate (NH4)2SO4. The MW of each fraction decreased consecutively with the concentration of (NH4)2SO4 at which it was precipitated. The MW distribution of each fraction was much narrower compared to the parent sample and is comparable to commercially available pullulan MW standards. To determine whether the fractionation process was separating sub-fractions of different structure, the original β-glucan sample and each fraction were hydrolyzed by a (1→3)(1→4)-D-β-glucan-4-glucanohydrolase (lichenase, E.C.3.2.1.73) and the liberated oligosaccharides were analyzed by high performance anion exchange chromatography. The analysis revealed no differences in oligosaccharide pattern (DP 2–9) derived from each fraction and the parent sample. In particular, the tri/tetra oligosaccharide ratio remained constant for all fractions, indicating no fractionation based on structural features had taken place. The effect of starting β-glucan concentration on the fractionation process was studied. The results showed that it was possible to achieve good separation at overlapping parameter c[η] lower than 3.5. Further increase in starting β-glucan concentration hindered clear separation of the fractions. Temperature also affected the fractionation efficiency. The higher the temperature, the lower the amount of (NH4)2SO4 that was necessary to precipitate the samples of same MW. A Mark Houwink relationship was derived from the measured MW and intrinsic viscosity for fractions from oat and barley, respectively.  相似文献   

16.
17.
The influence of high-intensity pulsed electric field (HIPEF) parameters, pulse frequency, pulse width and pulse polarity in strawberry juice lipoxygenase (LOX) and β-glucosidase (β-GLUC) was studied using a response surface methodology. The studied parameters affected on both residual enzymatic activities at unchanging electric field strength of 35 kV/cm and treatment time for 1000 μs. The contour plots showed a minimum defined space where residual activity of LOX remained at 65% and 70% in monopolar and bipolar mode, respectively. Low pulse frequencies (up to 61.6 Hz) in monopolar treatments as well as pulse frequencies and widths higher than 218 Hz and 5.4 μs in bipolar treatments did not have any effect on LOX inactivation. On the other hand, the higher the pulse frequency and pulse width, the higher the β-GLUC inactivation obtained. Moreover, when the HIPEF treatment was applied in monopolar mode, an enhancement in β-GLUC activity was observed in most of the experimental range. HIPEF treatments have demonstrated adequately that can reduce activity of enzymes that are involved in the formation of desirable flavor compounds, helping processors to obtain juices that keep their fresh flavor.

Industrial relevance

High-intensity pulsed electric fields (HIPEF) have proved to be effective in the interaction of microorganisms and enzymes in juices, maintaining their quality and freshness.HIPEF juice processing has demonstrated to have some advantages with regard to conventional thermal treatment. HIPEF treatments can reduce adequately enzymes that are involved in the formation of desirable flavor or color compounds. Thus, HIPEF technology can help processors to obtain juices that keep their fresh flavor by achieving optimal inactivation of related enzymes. This would prevent the product from undesirable off-flavor formation, which in turn would result in greater acceptability by consumers.  相似文献   

18.
The gelling properties of κ/ι hybrid carrageenans extracted from Portuguese seaweeds have been studied for various concentrations and shear conditions. The gel equilibrium storage modulus in 0.05 m KCl shows a power law dependence with the biopolymer concentration. Data analysis with a fibril model indicates that the fractal dimension of networking fibrils differ from the one displayed by pure κ‐ or ι‐carrageenan fibrils. κ/ι hybrid carrageenan gels show thermal hysteresis reminiscent from κ‐carrageenan gels under similar salt conditions. Application of steady shear rate during the cooling of a hot κ/ι hybrid carrageenan solution impedes the gel formation. However, a shear‐rate‐dependent gel structural rebuilding is observed after shear cessation. Shear‐processed gels are softer and show a depressed melting temperature when compared with gels formed under quiescent conditions. These experimental results are compared with data obtained with pure κ‐carrageenan and pure ι‐carrageenan solutions under similar steady shear and salt conditions.  相似文献   

19.
This work comprises the study of the thermal treatment of β-lg and its denaturation as a function of pH and ionic strength followed by differential scanning calorimetry. The concentration of protein was 14 (w/v)% in order to study the behaviour of highly concentrated β-lg solutions during heating. The denaturation temperature of β-lg was dependent on both pH and ionic strength, meaning that electrostatic interactions between protein monomers in the native state were important for the denaturation of β-lg. The thermograms from the calorimetric measurements also revealed that the quarternary structure of β-lg at pH-values close to the isoelectric point was influenced by the presence of salt and the nature of the salt (NaCl, KI and LiI). Small exotherms emerged in the thermograms at the low temperature side of the denaturation temperature for β-lg. The presences of these exotherms are probably caused by restructuring of the quarternary structure of native β-lg prior to denaturation, due to dissociation into smaller entities and possible also formation of a liquid crystalline-like structure in the highly concentrated protein solution. The present study provides a contribution to the understanding of the importance of the electrostatic interactions between native β-lg molecules and how different salts and ionic strengths affect the denaturation properties of the protein in concentrated systems.  相似文献   

20.
Gel formation by β-conglycinin and glycinin and their mixtures   总被引:1,自引:0,他引:1  
Gel formation and gel properties of β-conglycinin, glycinin and their mixtures were studied as a function of pH using small and large deformation rheology and differential scanning calorimetry. We conclude that heat denaturation is a prerequisite for gel formation. Gelation temperatures of β-conglycinin were lower than those of glycinin and more dependent on protein concentration. At pH 7.6, protein solutions gelled at a higher temperature than at pH 3.8.Glycinin gels were stiffer than β-conglycinin gels at the same pH and protein concentration, and fractured at a higher strain and stress. At pH 7.6, G′ is lower than at pH 3.8 for both proteins and the gels could be deformed to a larger extent. Based on the appearance of the gels (turbid at pH 7.6, white at pH 3.8) and the fracture properties, we conclude that different network structures are formed as a function of pH. The reason why glycinin gives a better gel than β-conglycinin is believed to be due to a difference in network structure as well as in strength of interaction between the protein molecules.Mixing of both soy proteins resulted in improved gelling properties at pH 3.8. The elastic modulus of the mixture was larger than the weighed sum of the separate contributions. Furthermore, mixing reduced the protein dispersability at pH 7.6. This strongly indicates the presence of an interaction between the proteins. Gels of the 1:1 mixture (pH 3.8) had a fracture stress and strain in between those of the gels of the separate proteins.  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