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 125 毫秒
1.
《Cereal Chemistry》2017,94(3):485-490
Proteins from wheat and egg are important for pound cake texture, but their exact role is insufficiently understood. A clear, analytical distinction between proteins from wheat flour, egg white, or egg yolk has been a main challenge. However, this can be addressed by using egg proteins carrying 15N. Therefore, egg white and yolk protein were enriched in 15N by mixing 15N‐labeled leucine into hen feed. Incorporation of egg and flour proteins in the protein network was monitored based on changes in their extractability during cake making. The relative contribution of different noncovalent and covalent bonds could be determined by using different extraction media. We for the first time distinguished between the contribution of egg white, egg yolk, and wheat protein in network formation during pound cake making. Our results show that during batter mixing hardly any intermolecular disulfide bonds are formed and that baking induces tremendous changes in protein extractability. A protein network based on both disulfide bonds and hydrophobic interactions is formed during baking. This covalent network includes almost all egg white protein and most of the yolk and wheat flour protein. The remaining protein fraction most probably lacks sulfhydryl groups and/or intramolecular disulfide bonds.  相似文献   

2.
Whey protein concentrate solutions (12% w/v, pH 6.65 +/- 0.05) were pressure treated at 800 MPa for 20-120 min and then examined using size exclusion chromatography (SEC), small deformation rheology, transmission electron microscopy, and various types of one-dimensional (1D) and two-dimensional (2D) polyacrylamide gel electrophoresis (PAGE). The pressure-treated samples showed a time-dependent loss of native whey proteins by SEC and 1D PAGE and a corresponding increase in non-native proteins and protein aggregates of different sizes. These aggregates altered the viscosity and opacity of the samples and were shown to be cross-linked by intermolecular disulfide bonds and by noncovalent interactions using 1D PAGE [alkaline (or native), sodium dodecyl sulfate (SDS), and SDS of reduced samples (SDS(R))] and 2D PAGE (native:SDS and SDS:SDS(R)). The sensitivity of the major whey proteins to pressure was in the order beta-lactoglobulin B (beta-LG B) > beta-LG A > bovine serum albumin (BSA) > alpha-lactalbumin (alpha-LA), and the large internal hydrophobic cavity of beta-LG may have been partially responsible for its sensitivity to high-pressure treatments. It seemed likely that, at 800 MPa, the formation of a beta-LG disulfide-bonded network preceded the formation of disulfide bonds between alpha-LA or BSA and beta-LG to form multiprotein aggregates, possibly because the disulfide bonds of alpha-LA and BSA are less exposed than those of beta-LG either during or after pressure treatment. It may be possible that intermolecular disulfide bond formation occurred at high pressure and that hydrophobic association became important after the high-pressure treatment.  相似文献   

3.
The interactions taking place in composite dough containing rice flour and soybean proteins (5% w/w) in the presence of transglutaminase, an enzyme with cross‐linking activity, were studied using different electrophoretic analyses. The interaction between rice proteins and soybean proteins was intensified by the formation of new intermolecular covalent bonds catalyzed by transglutaminase and the indirect formation of disulfide bonds among proteins. The main protein fractions involved in those interactions were both β‐conglycinin and glycinin of soybean and the glutelins of the rice flour, although albumins and globulins were also cross‐linked. The addition of soybean proteins to rice flour improves the amino acid balance and they also might play an important role on the rice dough properties because soybean proteins interact with rice proteins, yielding protein aggregates of high molecular weight.  相似文献   

4.
The disulfide bonds of beta-lactoglobulin (beta-lg) were modified by oxidative sulfitolysis to generate beta-lgSO(3). The native protein (beta-lg) and the modified protein (beta-lgSO(3)) were conjugated to activated polyethylene glycol (PEG) to generate beta-lgPEG and beta-lgSO(3)PEG, respectively. Oil-in-water (o/w) emulsions containing 1% beta-lg or beta-lg conjugates were prepared at pH 2.8, 5.0, and 7.0. Emulsion droplet diameters and zeta potentials were measured. For the same emulsifier, emulsion droplet diameters decreased when emulsion pH increased. Zeta potentials of emulsion droplets increased with pH for beta-lg and beta-lgSO(3). Zeta potentials of beta-lgPEG and beta-lgSO(3)PEG approached zero, suggesting that the protein molecule was covered by PEG chains. Accelerated and 7-day storage stabilities at 21 degrees C of the emulsions were monitored. The emulsifying activity index (EAI) of beta-lgPEG was not significantly different from the EAI of beta-lg. The EAI of beta-lg was enhanced following sulfitolysis of beta-lactoglobulin. The emulsifying activity increased more when the oxidatively modified protein was conjugated to polyethylene glycol. Emulsions made with beta-lgSO(3)PEG were more stable than emulsions made with beta-lg, beta-lgPEG, or beta-lgSO(3) under accelerated stability study and for 7 days at 21 degrees C. The stability of o/w emulsions stabilized with beta-lgSO(3)PEG increased because individual droplets were better protected, against protein bridging or coalescence, by the thick adsorbed protein-PEG layer.  相似文献   

5.
为提高酸性条件下大豆分离蛋白(soy protein isolates,SPI)的乳化性能,该文研究了物理-酶联合改性对SPI(pH值为4)的乳化性能影响,通过对比确定了物理-酶联合改性,即超声波-酶复合改性和挤压膨化-酶复合改性两种改性方法在酸性条件下的乳化性能效果最好;并通过对改性后 SPI(pH 值为4)进行溶解性、游离巯基、二硫键、粒径、扫描电镜(scanning electron microscope,SEM)和激光共焦扫描显微镜(confocal laser scanning microscopy,CLSM)分析,从蛋白结构变化上进一步揭示了乳化性能提高现象的原因。结果表明:超声波联合植酸酶-酸性蛋白酶改性的 SPI (Uphy-aci-SPI)的乳化活性(emulsifying activity index,EAI)为0.53 m2/g,比未改性SPI(0.18 m2/g)显著提高了196%(P<0.05),乳化稳定性(emulsifying stability index,ESI)为17 min,比未改性SPI(13.5 min)显著提高了25.9%(P<0.05);挤压膨化联合菠萝蛋白酶改性的SPI(Ebro-SPI)的EAI为0.46 m2/g,比未改性SPI显著增加了155%(P<0.05),ESI为17 min,比未改性SPI显著增加了25.9%(P<0.05)。在pH值为4的条件下对物理-酶联合改性的SPI的性质分析发现,物理-酶联合改性的SPI与未改性SPI相比,物理-酶联合改性的SPI的溶解性显著增加(P<0.05);物理-酶联合改性的SPI的乳状液平均粒径减小,CLSM观察乳状液中油与蛋白溶液稳定共融,改善了油滴之间的空间排斥力。物理-酶联合改性的SPI游离巯基的含量显著增加(P<0.05),二硫键含量显著降低(P<0.05)。SEM观察物理-酶联合改性的SPI为结构松散、破碎均一的微观结构。由此可见,乳化性能的提高是通过深层改变蛋白的结构来实现的。该研究可为探索提高酸性条件下SPI的乳化性能的方法提供理论依据。  相似文献   

6.
Cold gelation of whey proteins is a two-step process. First, protein aggregates are prepared by a heat treatment of a solution of native proteins in the absence of salt. Second, after cooling of the solution, gelation is induced by lowering the pH at ambient temperature. To demonstrate the additional formation of disulfide bonds during this second step, gelation of whey protein aggregates with and without a thiol-blocking treatment was studied. Modification of reactive thiols on the surface of the aggregates was carried out after the heat-treatment step. To exclude specific effects of the agent itself, different thiol-blocking agents were used. Dynamic light scattering and SDS-agarose gel electrophoresis were used to show that the size of the aggregates was not changed by this modification. The kinetics of gelation as determined by the development of pH and turbidity within the first 8 h of acidification were not affected by blocking thiol groups. During gelation, formation of large, covalently linked, aggregates occurred only in the case of unblocked WPI aggregates, which demonstrates that additional disulfide bonds were formed. Results of permeability and confocal scanning laser microscope measurements did not reveal any differences in the microstructure of networks prepared from treated or untreated whey protein aggregates. However, gel hardness was decreased 10-fold in gels prepared from blocked aggregates. Mixing different amounts of blocked and unblocked aggregates allowed gel hardness to be controlled. It is proposed that the initial microstructure of the gels is primarily determined by the acid-induced noncovalent interactions. The additional covalent disulfide bonds formed during gelation are involved in stabilizing the network and increase gel strength.  相似文献   

7.
Atomic force microscopy (AFM) has been used to study the noncovalent interactions of alkylated HMW subunit 1Dx5 and a M r 58,000 peptide derived from the central repetitive domain. Both protein and peptide align side‐by‐side to form fibrils, the HMW subunit forming a branched network, and the peptide forming linear rods. The N‐ and C‐terminal domains of the subunit would, therefore, appear to contain regions that interact through noncovalent interactions in the absence of disulfide bond formation. These regions may be of importance in facilitating disulfide bond formation during protein body development.  相似文献   

8.
Rice endosperm protein was modified to enhance solubility and emulsifying properties by controlled enzymatic hydrolysis. The optimum degree of hydrolysis (DH) was determined for acid, neutral, and alkaline type proteases. Solubility and emulsifying properties of the hydrolysates were compared and correlated with DH and surface hydrophobicity. DH was positively associated with solubility of resulting protein hydrolysate regardless of the hydrolyzing enzyme, but enzyme specificity and DH interactively determined the emulsifying properties of the protein hydrolysate. The optimum DH was 6–10% for good emulsifying properties of rice protein, depending on enzyme specificity. High hydrophobic and sulfhydryl disulfide (SH-SS) interactions contributed to protein insolubility even at high DH. The exposure of buried hydrophobic regions of protein that accompanied high-temperature enzyme inactivation promoted aggregation and cross-linking of partially hydrolyzed proteins, thus decreasing the solubility and emulsifying properties of the resulting hydrolysate. Due to the highly insoluble nature of rice protein, surface hydrophobicity was not a reliable indicator for predicting protein solubility and emulsifying properties. Solubility and molecular flexibility are the essential factors in achieving good emulsifying properties of rice endosperm protein isolates.  相似文献   

9.
为提高米糠谷蛋白功能性质,本研究对米糠谷蛋白与β-环状糊精进行复合热处理(温度60、70、80、90、99℃,时间40、80、120、160、200 min),分析复合聚集体的浊度、接枝度、乳化性质及结构特性等,探究米糠谷蛋白与β-环状糊精复合热聚集行为。结果表明,米糠谷蛋白与β-环状糊精在90℃条件下加热复合160 min时,复合聚集体乳化活性指数达到最大,与天然米糠谷蛋白相比提高了2.39倍;在80℃条件下加热复合200 min时,复合聚集体乳化稳定性指数最大,与天然米糠谷蛋白相比提高了2.39倍。复合物在80℃条件下受热后,米糠谷蛋白与β-环状糊精结合生成较大颗粒的聚集体;复合物中米糠谷蛋白肽链结构打开,游离巯基含量增加,二硫键断开,疏水基团暴露,β-折叠向α-螺旋和β-转角转化,以共价键的形式形成分子间氢键,使得复合聚集体分子更好地结合到油水的界面,复合聚集体乳化活性和稳定性显著提高(P<0.05)。本研究结果为后续米糠蛋白质功能性质的改善及米糠蛋白质的深加工提供了理论依据。  相似文献   

10.
Dough development using sheeting and mechanical dough development (MDD) were compared with respect to the effect the mixing method had on the molecular size distribution and degree of protein thiol exposure of the aggregated glutenin proteins. Although sheeting imparts a lower rate of work input on doughs than does MDD mixing, changes in protein aggregation patterns during mixing were similar for both methods of dough development, indicating that protein disaggregation was important in the process of dough development. In both systems, a reduced rate of change in the protein aggregation patterns was associated with optimum dough development. The MDD mixing was characterized by increasing exposure of the thiol groups on the SDS‐insoluble glutenin during mixing while the sheeting process resulted in fewer exposed thiol groups on both SDS‐soluble and SDS‐insoluble glutenin proteins. This suggested that disulfide bond rupture may not be a required process in dough development and that high effective stresses per se may not be required to develop doughs. This is consistent with a model for dough development that does not require extensive covalent bond rupture but instead involves mainly rupture and reformation of noncovalent interactions such as hydrophobic bonds and hydrogen bonds between protein chains.  相似文献   

11.
《Cereal Chemistry》2017,94(4):733-739
The present study tried to phosphorylate rice protein (RP), a known insoluble food ingredient, and determine the improvement of its solubility. RP was allowed to react with sodium trimetaphosphate (STMP) at pH 11.5 and 35°C, and the results indicated that 20.6% of the RP seryl residues were phosphorylated. Interestingly, the solubility of phosphorylated RP (2.6%) was not improved compared with that of RP (2.5%) at pH 7. The involvement of hydrophobic interactions and disulfide bonds in phosphorylated RP solubility was further evaluated. The phosphorylation of RP in the presence of urea as a chaotropic agent for weakening the hydrophobic effect resulted in 22.0% phosphoseryl residues but still did not increase RP solubility. The reduction of RP disulfide bonds prior to phosphorylation resulted in 31.3% phosphoseryl residues and increased RP solubility to 8.3% at pH 7, indicating that disulfide bonds within RP could be responsible for the failure to increase its solubility after phosphorylation.  相似文献   

12.
Chemical treatments with a number of low‐toxicity or nontoxic reagents were applied to corn slurry to investigate the disruption or weakening of common bonding forces between corn starch and proteins, such as hydrogen bonds, disulfide bonds, electrostatic interactions, or combinations thereof, to improve the corn starch isolation process. Starch and proteins could be easily separated by disrupting disulfide bonds with 1% l ‐cysteine (w/v). The most effective reagents for hydrogen bonds and electrostatic interactions were 3M urea and pH 7.5 separately. The sequence treatment of hydrogen bonds, disulfide bonds, and electrostatic interactions (namely, sequence treatment of 1M urea, 1% l ‐cysteine, and pH 7.5) led to the highest amount of starch in corn slurry and facilitated the corn starch isolation.  相似文献   

13.
Emulsifying properties of native and chemically modified soy glycinins were studied. The influence of ionic strength, protein sample composition and concentration, and assay conditions on the flocculation-creaming process and coalescence resistance was analyzed. Differences in these emulsifying properties were exhibited by native glycinins, which have a variable content of 4S, 11S, and 15S forms. Structure and functionality of native glycinin were modified by means of combined treatments: mild acidic treatments without heating or with heating at variable time and with or without disulfide bonds reduction. Modified glycinins presented different degrees of deamidation, surface hydrophobicity, and molecular mass. A slight enhancement of emulsifying stability at moderated deamidation degrees was observed. In different protein samples, a positive relationship between the flocculation-creaming rate constant and equilibrium oil volume fraction of emulsions with surface hydrophobicity was detected. A remarkable difference was observed between reduced and nonreduced samples, mainly with respect to behavior at low or high ionic strength.  相似文献   

14.
Kunitz trypsin inhibitor (KTI) and Bowman-Birk inhibitor (BBI) have trypsin inhibitor activities (TIA), which could cause pancreatic disease if at a high level. It is not clear why some KTI and BBI lose TIA and some does not in the soymilk processing. This would be examined in this study. TIA assay showed residual TIA was decreased with elevated temperature and TIA was decreased quickly in the beginning and then slowly in boiling water bath. Interestingly, ultracentrifugation showed low residual TIA soymilk had more precipitate than high residual TIA soymilk and soymilk TIA loss had a high correlation coefficient (R(2) > 0.9) with precipitate amount. In addition, the TIAs of floating, supernatant, and precipitate obtained by ultracentrifugation were assayed and >80% residual TIA was concentrated in the supernatant. Tricine-SDS-PAGE showed KTI in supernatant was mainly a noncovalent bound form which might exist as itself and/or incorporated into a small protein aggregate, while KTI in precipitate was incorporated into a protein aggregate by disulfide and/or noncovalent bonds. Chymotrypsin inhibitor activity (CIA) assay showed about 89% of the original CIA remained after 100 °C for 15 min. Ultracentrifugation showed that >90% residual CIA was concentrated in supernatant. Tricine-SDS-PAGE showed soymilk (100 °C, 15 min) BBI mainly existed in supernatant but not in precipitate. It was considered that BBI tended to exist as itself with its natural conformation. Thus, it was suggested residual TIA was mainly from the free BBI and TIA inactivation was mainly from KTI incorporation into protein aggregate. This study is meaningful for a new strategy for low TIA soymilk manufacture based on the consideration of promoting protein aggregate formation.  相似文献   

15.
Intermolecular disulfide bond formation in pure beta-lactoglobulin (beta-Lg) B and in its 1:1 mixture with alpha-lactalbumin (alpha-La), heated at 85 degrees C for 10 min in solutions of low and high (100 mM NaCl) ionic strength and pH 6.0, was studied by reverse-phase HPLC and MALDI-TOF mass spectrometry. Disulfide bonding between beta-Lg monomers was more extensive than reported in the literature for a temperature of 68.5 degrees C, including formation of trimers connected by two of the three adjacent cysteines, C106/C119/C121. The participation of the different thiol groups in disulfide bonds appeared to depend on their location in the native structure, with surface-located cysteines more involved than internally located ones. This also applied to alpha-La-beta-Lg interactions, where the predominant participants were the surface-located alphaC111, alphaC120, alphaC61, and alphaC6. The least active participant was alphaC28, suggesting that it becomes sterically inaccessible during unfolding of the protein. High ionic strength apparently promoted disulfide bonding. The order of cysteine participation at the high ionic strength was similar to that at low ionic strength, with fewer native-location bonds observed and a lower activity of some groups, such as beta-C106/C119/C121 and alphaC61.  相似文献   

16.
Thaumatin, a sweet protein that contains no cysteine residues and eight intramolecular disulfide bonds, aggregates upon heating at pH 7.0 above 70 degrees C, and its sweetness thereby disappears. The aggregate can be solubilized by heating in the presence of both thiol reducing reagent and SDS. This molecular aggregation depended on the protein concentration during heating and was suppressed by the addition of N-ethylmaleimide or iodoacetamide, indicating a thiol-catalyzed disulfide interchange reaction between heat-denatured molecules. An amino acid analysis of the aggregates suggested that the cysteine and lysine residues were reduced, and the formation of a cysteine residue and a lysinoalanine residue was confirmed. The reduction and formation of these residues stoichiometrically satisfied the beta-elimination of a cystine residue. The disulfide interchange reaction was catalyzed by cysteine; that is, a free sulfhydryl residue was formed via beta-elimination of a disulfide bond. Intermolecular disulfide bonds were probably formed between thaumatin molecules upon heating at pH 7.0, which led to the aggregation of thaumatin molecules.  相似文献   

17.
Napin from mustard (Brassica juncea L.) is a seed storage protein consisting of two subunits linked through disulfide bonds and is predominantly helical in nature. Resistance to trypsin digestion and allergenicity limit its food applications. The role of disulfide linkages, electrostatic as well as hydrophobic interactions, in napin stability have been investigated through spectroscopic methods, employing different fluorescent probes and additives. The subunits are hydrophilic in nature and possess extended structure. With the addition of 0.5 M NaCl, the surface hydrophobicity of napin decreases, whereas the helical content increases by 25%. In the presence of NaCl, emission maximum shifts toward shorter wavelength and the Stern-Volmer constant decreases from 6.5 to 3.4 M-1, indicating compaction of napin. Na2SO4 has no significant effect on the structure due to the lack of a hydrophobic core. In the presence of monohydric alcohols and trifluoroethanol, there is an increase in ordered structure. These studies indicate that the structure of napin, which is hydrophilic in nature, is stabilized by electrostatic interactions, in addition to disulfide linkages.  相似文献   

18.
Soy protein, mixed with gluten and starch, was extruded into fibrous meat analogues under high-moisture and high-temperature conditions. The protein solubility of samples collected at different extruder zones and extrudates made with different moistures was determined by 11 extraction solutions consisting of 6 selective reagents and their combinations: phosphate salts, urea, DTT, thiourea, Triton X-100, and CHAPS. Protein solubility by most extractants showed decreasing patterns as the material passed through the extruder, but the solution containing all 6 reagents, known as isoelectric focus (IEF) buffer, solubilized the highest levels and equal amounts of proteins in all samples, indicating that there are no other covalent bonds involved besides disulfide bonds. With regard to relative importance between disulfide bonds and non-covalent interactions, different conclusions could be made from protein solubility patterns, depending on the type of extracting systems and a baseline used for comparison. The observation points out pitfalls and limitation of current protein solubility methodology and explains why controversy exists in the literature. Using the IEF buffer system with omission of one or more selective reagents is considered to be the right methodology to conduct protein solubility study and thus recommended. Results obtained with this system indicate that disulfide bonding plays a more important role than non-covalent bonds in not only holding the rigid structure of extrudates but also forming fibrous texture. The sharpest decrease in protein solubility occurred when the mix passed through the intermediate section of the extruder barrel, indicating formation of new disulfide bonds during the stage of dramatic increase in both temperature and moisture. After this stage, although the physical form of the product might undergo change and fiber formation might occur as it passed through the cooling die, the chemical nature of the product did not change significantly.  相似文献   

19.
An attempt was made to evaluate gluten structural changes in refined and whole wheat pasta from hard white winter wheat to elucidate the impact of whole wheat components on the formation and structure of the gluten network in pasta. Attenuated total reflectance–FTIR spectroscopy was used to track gluten secondary structure through most of the major steps in pasta processing: raw material, mixing, drying, and cooking. Protein solubility, accessible thiols, and SDS‐PAGE data were also collected to provide additional information on the nature of protein interactions and network composition. Few secondary structural differences were observed between refined and whole wheat flours from hard white wheat. However, mixing induced a significant shift to β‐sheet structures in refined dough that was not equally matched by whole wheat dough. Drying under both high temperature, short time (HT) and low temperature, long time (LT) conditions resulted in a reversion to structural distributions similar to those for flour in both pastas. However, greater protein denaturation in HT samples was indicated by lower protein solubility also in the presence of denaturants and disulfide reducing agents. Cooking generated a substantial increase in β‐sheet structures for both pasta systems. This structure was greatest in refined and LT samples. Thiol accessibility data indicate the presence of a highly aggregated, compact gluten network in refined pasta, mostly driven by hydrophobic association. Conversely, the network in whole wheat pasta was more loosely associated and dependent on disulfide bonding, both of which fit well with the secondary structural data.  相似文献   

20.
The amino acid composition and physicochemical and functional properties of hemp (Cannabis sativa L.) protein isolate (HPI) were evaluated and compared with those of soy protein isolate (SPI). Edestin, a kind of hexameric legumin, was the major protein component. HPI had similar or higher levels of essential amino acids (except lysine), in comparison to those amino acids of SPI. The essential amino acids in HPI (except lysine and sulfur-containing amino acids) are sufficient for the FAO/WHO suggested requirements for 2-5 year old children. The protein solubility (PS) of HPI was lower than that of SPI at pH less than 8.0 but similar at above pH 8.0. HPI contained much higher free sulfhydryl (SH) content than SPI. Differential scanning calorimetry analysis showed that HPI had only one endothermic peak with denaturation temperature (T(d)) of about 95.0 degrees C, attributed to the edestin component. The T(d) of the endotherm was nearly unaffected by 20-40 mM sodium dodecyl sulfate but significantly decreased by 20 mM dithiothreitol (P < 0.05). The emulsifying activity index, emulsion stability index, and water-holding capacity of HPI were much lower than those of SPI, and the fat adsorption capacity was similar. The data suggest that HPI can be used as a valuable source of nutrition for infants and children but has poor functional properties when compared with SPI. The poor functional properties of HPI have been largely attributed to the formation of covalent disulfide bonds between individual proteins and subsequent aggregation at neutral or acidic pH, due to its high free sulfhydryl content from sulfur-containing amino acids.  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