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.
Physicochemical and rheological properties of myosin from Atlantic cod and burbot during acid-induced gelation at room temperature (22–23 °C) by d-gluconic acid-δ-lactone (GDL) were monitored. Turbidity and particle size of both myosins increased and salt soluble content decreased when pH decreased, suggesting the formation of protein aggregates caused by acidification. The formation of disulphide bonds in myosin gelation was induced by acid. Ca2+-ATPase activity of myosin decreased (p < 0.05), while surface hydrophobicity increased during acidification (p < 0.05). Furthermore, the decreases in maximum transition temperature (Tmax) and the denaturation enthalpies (ΔH) were found in both myosins. During acidification, the increases in storage modulus (G′) and loss modulus (G″) of myosin were observed (p < 0.05), revealing the formation of elastic gel matrix. Thus, gelation of myosin from Atlantic cod and burbot could take place under acidic pH via denaturation and aggregation. However, myosin from Atlantic cod was generally more favourable to gelation than was burbot myosin.  相似文献   

2.
Gelation is one of the most important properties of plant proteins. In this paper, a low denaturation salt extraction method was used to extract pea (Pisum sativum L.) protein isolate from commercial pea flour. The gelation properties of this isolate were examined and compared to commercial products. The pea protein isolate followed the three-step process of gelation that is generally accepted for heat-induced gelation of globular proteins. The minimum gelation concentration of salt-extracted pea protein isolate (PPIs) was 5.5% while that of commercial pea protein isolate (PPIc) was 14.5%. The gelling point was in the range of 82–86 °C for 14.5% PPIs, 0.3 M NaCl at natural pH (5.65). With increasing heating rate, the gelling point tended to increase. Higher heating and cooling rates resulted in decreased final G′ (storage modulus) and G″ (loss modulus) values, indicative of decreased gel strength. A higher protein concentration resulted in higher G′ and G″ values and it was found that there was a power law relationship between protein concentration and G′ and G″. Tan delta (δ) values decreased with increasing protein concentration and at concentrations of 5.5% and above, tan δ remained constant which means the critical concentration for gel formation was 5.5%. The values of G′ and G″ for PPIs were greater than those of PPIc, and tan δ of PPIs was smaller, indicative of a stronger gel network. DSC data showed that PPIc had undergone denaturation whereas PPIs had not (ΔH = 15.81 J/g protein). Although rheometer data showed that the final G′ value of commercial soy protein isolate (SPIc) was smaller than that of PPIs, the gel prepared with SPIc was visually stronger than that of PPIs. The rheological data obtained with small amplitude oscillatory testing was not consistent with the actual observations. Overall, the low degree of denaturation of the PPIs resulted in a stronger gel than that of PPIc making the PPIs a more attractive food ingredient.  相似文献   

3.
Gelation properties of chicken myofibrillar protein isolate (MPI) and the effect of microbial transglutaminase (MTG) were studied using a dynamic oscillatory rheometer and a texture analyzer. Final heating temperature had a great impact on gel stiffness and the maximum gel stiffness was obtained at 95 °C. pH and ionic strength also influenced gel stiffness and the maximum gel stiffness was achieved at pH 6, 0.9 M NaCl; however, less stiff gels were formed in 0.6 and 1.2 M NaCl. In the MPI concentration range of ∼0.5-5%, a positive correlation was observed between gel stiffness or gel peak force and MPI concentration. When MTG was included at levels of ∼0 to 12-15 U, positive linear relations were found between gel stiffness or peak force and MTG levels. However, negative correlations for these parameters were observed at higher MTG concentrations. When MTG level was greater than 15 U, gel stiffness or peak force tended to decrease. The improvement in gel strength or gel peak force for the MPI with inclusion of MTG suggested that some ε (γ-glutamyl) lysine (G-L) crosslinking occurred among myofibrillar molecules. Thus, MTG is useful in improving gelation properties of heat-induced MPI gel and provides new opportunities to expand the utilization of low value meat in muscle foods.  相似文献   

4.
The effects of phosphorylation with sodium trimetaphosphate (STMP) on functional and physicochemical properties of pea vicilin, as probed by high hydrostatic pressure and chemical denaturation were evaluated. The isoelectric point of unmodified and phosphorylated vicilin decreased in the presence of 0.5 M NaCl, resulting in a decrease of the solubility at pH 1.0. The gelation capacity of unmodified vicilin in the presence of NaCl decreased approximately 80% when compared with unmodified vicilin without NaCl. Increasing pressure from 0.1 MPa (atmospheric pressure) to 240 MPa significantly decreased the solubility of vicilin phosphorylated with 4% STMP at pH 1.0 and 4.0 by about 30%. Pressure had no effect on solubility of native vicilin. Pressure treatment at 240 MPa improved the gelation capacity of vicilin phosphorylated with 1% STMP. Glycerol decreased the gelation capacity of vicilin and its solubility in the acidic pH range.  相似文献   

5.
This study investigated the effects of sodium alginate (SA) and glucono-δ-lactone (GdL) levels on the cold-set gelation of porcine myofibrillar protein (MP) at different salt concentrations. The addition of GdL at 0.1 M salt concentration had no effects on cold-set MP gelation, and GdL level higher than 1% was necessary at 0.3 M salt concentration to form a MP gel. The SA system with 0.5% GdL showed an effect on MP gelation at 0.1 M salt, while SA with 1.5% GdL could be form a cold-set MP gel at 0.3 M salt level. The endothermic peaks were reduced or completely disappeared by the addition of GdL more than 1%. These results indicated that the SA was favorable in forming a cold-set MP gel under low salt concentrations (<0.1 M), and that the SA could be applicable at higher salt concentrations by combining with the GdL level more than 1% (>1%).  相似文献   

6.
Cony Gauche 《LWT》2010,43(2):214-219
The effect of thermal treatment of whey proteins (50 g/100 g) at 75, 80, 85, 90 and 95 °C before enzymatic treatment with microbial transglutaminase was evaluated by rheological measurements. Thermal denaturation of whey proteins was determined by differential scanning calorimetry (DSC); and the gel point and turbidity were also evaluated after addition of transglutaminase at different pH values in protein solutions. A significant (p < 0.05) interaction was observed between transglutaminase and thermal treatments. At temperatures higher than 85 °C the apparent viscosity measurements of whey protein solutions with transglutaminase were significantly higher than those of the control samples. DSC analysis showed that thermal denaturation occurred at temperatures close to 82 °C, and the enzymatic reaction was enhanced at higher temperatures. The gel point of whey proteins decreased with transglutaminase addition. This decrease became greater as a function of reaction time due to the formation of high weight protein polymers catalyzed by transglutaminase, which was also observed in the turbidity analysis.  相似文献   

7.
AAPH-derived (2,2′-azobis (2-amidinopropane) dihydrochloride) peroxyl radicals were selected as representative free radicals of lipid peroxidation to investigate the effects of oxidative modifications on isolated porcine myofibrillar protein structures as well as their rheological and gelling properties. Incubation of myofibrillar protein with increasing concentrations of AAPH resulted in a gradual increase (p < 0.05) in carbonyl content and SH → S–S conversion. Results from SDS-PAGE indicated that medium (~ 1 mM) and relatively high (> 3 mM) concentrations of AAPH induced aggregation of myosin and denaturation of myosin, troponin and tropomyosin, respectively. These structural changes resulted in changes on gelation of myofibrillar protein. Low level protein oxidation (AAPH ≤ 0.5 mM) had no remarkable effect (p > 0.05) on the viscoelastic pattern of myofibrillar protein gelation. Moderate oxidative modification (AAPH ~ 1 mM) enhanced the water-holding capacity (WHC) and texture properties of gels, while further oxidation (AAPH > 3 mM) significantly reduced the gel quality.  相似文献   

8.
Gels from silver carp (Hypophthalmichthys molitrix) surimi were obtained using microwave (MW) heating (15 W/g power intensity for 20–80 s) at different levels of salt (0 g/100 g, 1 g/100 g, or 2 g/100 g). And the gel heated by MW was compared with the gel obtained by conventional water-bath heating (85 °C for 30 min). The gel strength increased when the salt level was increased. The mechanical and functional properties of non-salted, low-salt and regular-salt products were improved by MW heating for 60 s and 80 s, significantly (p < 0.05), except for the cook loss. The content of TCA-soluble peptides indicated that the MW heating inhibited the autolysis of proteins significantly (p < 0.05) during gelling. The SDS-PAGE and total content of –SH group proved that MW enhanced the cross-linking of proteins effectively through disulphide bonds and non-disulphide covalent bonds. The microstructure of the samples revealed that a fine compact network, with particles of protein aggregates, was formed in the low-salt gels (1 g/100 g) heated by MW for 60 s. All of these properties might be responsible for the formation of a superior textural low-salt gel induced by MW.  相似文献   

9.
The molecular forces involved in the gelation of heat-induced pea protein gel were studied by monitoring changes in gelation properties in the presence of different chemicals. At 0.3 M concentration, sodium thiocyanate (NaSCN) and sodium chloride (NaCl) showed more chaotropic characteristic and enhanced the gel stiffness, whereas sodium sulfate (Na2SO4) and sodium acetate (CH3COONa) stabilized protein structure as noted by increasing denaturation temperatures (Td) resulting in reduced storage moduli (G′). To determine the involvement of non-covalent bonds in pea protein gelation, guanidine hydrochloride (GuHCl), propylene glycol (PG), and urea were employed. The significant decrease in G′ of pea protein gels with the addition of 3 M GuHCl and 5 M urea indicated that hydrophobic interactions and hydrogen bonds are probably involved in pea protein gel formation. The increase in G′ with increasing PG concentration (5–20%), demonstrated hydrogen bonds and electrostatic interaction involvement. No significant influence was observed on G′ with addition of different concentrations of β-mercaptoethanol (2-ME), low levels of dithiothreitol (DTT), and up to 25 mM N-ethylmaleimide (NEM), which indicated that disulfide bonds are not required for gel formation, but data at higher DTT and NEM concentrations and slow cooling rates showed a minor contribution by disulfide bonds. Reheating and recooling demonstrated that gel strengthening during the cooling phase was thermally reversible but not all the hydrogen bonds disrupted in the reheating stage were recovered when recooled.  相似文献   

10.
The gelation characteristics of myofibrillar proteins are indicative of meat product texture. Defining the performance of myofibrillar proteins during gelation is beneficial in maintaining quality and developing processed meat products and processes. This study investigates the impact of pH on viscoelastic properties of porcine myofibrillar proteins prepared from different muscles (semimembranosus (SM), longissimus dorsi (LD) and psoas major (PM)) during heat-induced gelation. Dynamic rheological properties were measured while heating at 1 °C/min from 20 to 85 °C, followed by a holding phase at 85 °C for 3 min and a cooling phase from 85 to 5 °C at a rate of 5 °C/min. Storage modulus (G′, the elastic response of the gelling material) increased as gel formation occurred, but decreased after reaching the temperature of myosin denaturation (52 °C) until approximately 60 °C when the gel strength increased again. This resulted in a peak and depression in the thermogram. Following 60 °C, the treatments maintained observed trends in gel strength, showing SM myofibrils produced the strongest gels. Myofibrillar protein from SM and PM formed stronger gels at pH 6.0 than at pH 6.5. Differences may be attributed to subtle variations in their protein profile related to muscle type or postmortem metabolism. Significant correlations were determined between G′ at 57, 72, 85 and 5 °C, indicating that changes affecting gel strength took effect prior to 57 °C. Muscle type was found to influence water-holding capacity to a greater degree than pH.  相似文献   

11.
The objective of this work was to compare the gelation of soymilk particles induced by the acidification of a commercial starter culture with that resulting by addition of glucono-δ-lactone (GDL). Structure formation was followed using rheology, and the microstructure was observed by confocal microscopy. Acidification of lactic acid bacteria resulted in a higher gelation pH (pH 6.29 ± 0.05) compared to that of a gel induced by GDL (pH 5.9 ± 0.04). This difference was attributed to the longer time available for rearrangements of the soymilk particles in soymilk with starter cultures compared to the fast acidification by GDL. In spite of the earlier gelation pH, there were no observed differences in the final gel stiffness measured at pH 5.1, the value of tan δ, the frequency dependence and the linear viscoelastic range of the gels measured at the final pH. Microstructural observations also showed a similar protein network structure.  相似文献   

12.
The current study investigated the effect of a neutral polysaccharide, konjac glucomannan, on the heat-induced gelation of whey protein isolate (WPI) at pH 7. Oscillatory rheology (1 rad/s; 0.5% strain), differential scanning calorimetry and confocal laser scanning microscopy were used to investigate the effect of addition of konjac in the range 0-0.5% w/w, on the thermal gelation properties of WPI. The minimum gelling concentration for WPI samples was 11% w/w; the concentration was therefore held constant at this value. Gelation of WPI was induced by heating the samples from 20 to 80 °C, holding at 80 °C for 30 min, cooling to 20 °C, and holding at 20 °C for a further 30 min. On incorporation of increasing concentrations of konjac the gelation time decreased, while the storage modulus (G′) of the mixed gel systems increased to ∼1450 Pa for 11% w/w WPI containing 0.5% w/w konjac gels, compared to 15 Pa for 11% w/w WPI gels (no konjac). This increase in gel strength was attributed to segregative interactions between denatured whey proteins and konjac glucomannan.  相似文献   

13.
Protein isolates and concentrates were obtained from defatted cashew nut powder by two methods: alkaline extraction-isoelectric precipitation (IP) and alkaline extraction-methanol precipitation (MP). The functional properties of cashew nut protein isolates, concentrates and powder were significantly different (p < 0.05). Cashew nut protein isolate (CNPI) had higher water and oil absorption capacities (2.20 ml/g and 4.42 ml/g, respectively), emulsifying stability index (447%), foam capacity and stability (45% and 55%, respectively), and least gelation capacity (13.5%) than cashew nut protein concentrate (CNPC), which was also higher than that of defatted cashew nut powder (DCNP). However, emulsifying activity index (12.45%) and bulk density (0.31) of CNPI were lower than that of CNPC, which were also lower than that of DCNP. The water solubility of CNPI (95%) and CNPC (95%) was not significantly different (p > 0.05) among the samples, but was significantly different (p < 0.05) from that of DCNP (75%). The CNPI, CNPC and DCNP showed decreasing solubility with decreasing pH, with the minimum solubility being observed at a pH range of 4.0–4.5, confirming the isoelectric point of cashew proteins. However, higher water solubility, emulsifying activity, and foaming property were observed at an alkaline pH than at an acidic pH in all samples.  相似文献   

14.
单宏 《食品科学》2015,36(5):72-76
采用一种低变性的提取方法--盐法提取豌豆分离蛋白并使用流变仪测定其流变特性,结果表明:盐法豌豆分离蛋白最低凝胶形成质量浓度是5.5 g/100 mL,豌豆分离蛋白凝胶点与蛋白质量浓度无相关性。豌豆分离蛋白凝胶点随加热速率增加呈上升趋势。较高的加热和冷却速率能够增加反应终点储能模量(G’)和耗能模量(G”)值,因而降低凝胶强度。盐法豌豆分离蛋白凝胶韧性随蛋白质量浓度的增加而增大,且蛋白质量浓度与G’、G”值之间存在乘幂规律。tanδ值随蛋白质量浓度升高而降低,当蛋白质量浓度高于5 g/100 mL时,弱凝胶开始形成;当蛋白质量浓度高于7 g/100 mL时,tanδ值几乎保持恒定,表明凝胶弹性在此范围内基本恒定。比较盐法和商业豌豆分离蛋白凝胶性质后发现:盐法豌豆分离蛋白具有更好的凝胶性。  相似文献   

15.
Chemical and biochemical changes of aquacultured hybrid catfish fillet (Clarias macrocephalus × Clarias gariepinus) and its gel-forming ability as affected by age and sex of fish along with storage time were investigated. Fillets were stored at 4 °C for 0, 3, 6, 9, 12 and 15 days. There was no significant effect of sex and age of fish as well as storage time on fat, moisture and ash contents (P > 0.05). The total protein, water soluble protein, and salt soluble protein contents of the fillets significantly decreased with storage time (P < 0.05). On the other hand, pH, total volatile base nitrogen (TVB-N) and autolytic degradation products (ADP) increased as storage time continued (P < 0.05). Decreases in Ca2+-ATPase activity and gel properties were observed as storage time increased. However, there was no significant effect of either sex or age of fish on textural properties of gel (P > 0.05). Hybrid catfish fillet stored at 4 °C should be processed within 6 days.  相似文献   

16.
Fish gelatin extraction from wastes of fish Herring species (Tenualosa ilisha) was carried out by a series of pretreatment with 0.2 M Ca(OH)2 followed by 0.1 M citric acid and final water extraction at 50 °C for 3 h. The resulting fish gelatin preparation was evaluated for its dynamic viscoelastic properties, gelling and melting temperatures and gel strength. The gelling and melting temperatures of gelatin samples (at 6.67%, w/v) were obtained from differential scanning calorimetry and rheological studies. The melting temperature of extracted fish gelatin (EFG) obtained ranged from 16.2 to 16.7 °C compared to that of commercial fish gelatin gel (CFG), from 23.7 to 25.6 °C and halal bovine gelatin (HBG), from 26.5 to 28.7 °C. On the other hand, gelling temperatures of EFG, CFG and HBG ranged from 5.1 to 5.2 °C, 11.9 to 17.46 °C, and 12.6 to 19.33 °C, respectively. EFG gave gels with a considerably lower G′ values than CFG and HBG. The bloom strength of EFG gels at 6.67% (w/v) was 69.03 g which was much lower than HBG (336.2 g) and CFG (435.9 g). Enzyme transglutaminase was added in the amounts of 0.5, 1.0, 3.0 and 5.0 mg/g gelatin to modify the gel properties of the extracted fish gelatin. The modified EFG gels obtained had higher gel strengths of 101.1 g and 90.56 g with added transglutaminase of 1.0 and 3.0 mg/g, respectively. However with addition of 5.0 mg/g enzyme the gel strength increased only up to 75.06 g. SDS-PAGE of extracted gelatin gel showed protein band intensities for α1-chains and 53 kDa but in gels added with higher concentration of transglutaminase, these protein band intensities seemed to disappear.  相似文献   

17.
Physicochemical changes in surimi with salt substitute   总被引:2,自引:0,他引:2  
Protein endothermic transitions (thermal denaturation), rheological properties (protein gelation), and fundamental texture properties (shear stress and strain at mechanical fracture) of Alaska pollock surimi gels made with 0 (control), 1, 2, and 3 g/100 g of salt (NaCl) were determined and compared with equal molar concentration of salt substitute. Salt and salt substitute shifted the onset of myosin transition to higher temperature and resulted in larger myosin peaks (i.e., transition enthalpy). Endothermic transitions showed similar trends to rheological properties. The elastic modulus (G′) increased when salt or salt substitute was added to surimi, except at the highest concentration of salt and salt substitute. Salt and salt substitute also induced the onset of protein gelation (i.e., as measured by significant increase of G′) at lower temperature. Surimi gels with salt substitute and salt at equal molar concentrations had similar texture properties (shear stress and strain). Based on the present study, salt substitute can be used in the development of low-sodium surimi seafood products without significant change in gelation and texture.  相似文献   

18.
Physicochemical and gelation properties of surimi prepared from three species of mackerel were investigated. The highest whiteness with the lowest redness index corresponding to the lowest myoglobin content especially its oxidised form, metmyoglobin, was found in short-bodied mackerel (Rastrelliger brachysoma) surimi (p < 0.05). Frigate mackerel (Auxis thazard) surimi contained the highest lipid content (p < 0.05). The pH of all surimi was in the range of 6.58–6.80. The highest sulfhydryl group and Ca2+-ATPase activity was found in natural actomyosin extracted from short-bodied mackerel surimi (p < 0.05). The highest TCA-soluble peptide content was found in frigate mackerel surimi gels (p < 0.05). Kamaboko gel of short-bodied mackerel surimi exhibited the highest breaking force with the lowest expressible drip (p < 0.05). Heating regime had no effect on deformability of gels from Indian mackerel (Rastrelliger kanagurta) and short-bodied mackerel but not for frigate mackerel. The highest metmyoglobin content with the lowest whiteness was found in frigate mackerel surimi gel (p < 0.05). Therefore, short-bodied mackerel was the best suited for the production of surimi with superior functional attributes including whiteness and gel-forming ability.  相似文献   

19.
Cod (Gadus Morhua) muscle proteins were solubilized using alkaline treatment of the muscle. Solutions of similar protein composition were obtained between pH 10.5 and 12.0, however, pH > 11 was required for optimal yield. Addition of salt (up to 0.25 M NaCl) did not affect protein yield or composition. Light scattering showed that a significant fraction of the proteins was present as large self similar and flexible aggregates. When the pH was decreased below 10, gelation was observed below a critical temperature of about 25 °C, which could be reversed by heating. Slow irreversible aggregation was also observed leading to coarsening and syneresis of the gels or precipitation at higher temperatures. The rate of irreversible aggregation increased with decreasing pH and was fast below pH 8. Homogeneous thermo-reversible self supporting gels that were stable for a period of days could be prepared without heating at a narrow pH range between 8.5 and 9.5.  相似文献   

20.
Thermal aggregation of duck egg white solution (1 mg protein/ml, pH 7) was monitored in the presence of different NaCl concentrations (0–6%, w/w) across the temperature range of 20–90 °C. Duck egg white solution exhibited higher turbidity with coincidental increases in surface hydrophobicity and decreases in sulfhydryl group content as temperatures increased from 70 to 90 °C (p < 0.05). As NaCl concentration increased, the negative charge decreased, with coincidental increases in particle size of aggregate after heating to 90 °C. As visualised by confocal laser scanning microscopy, larger clusters of protein aggregates were observed with increasing NaCl concentrations. Major duck egg white protein with molecular mass of 45 kDa disappeared in the presence of 2–6% NaCl after heating above 80 °C, regardless of concentrations. Therefore, NaCl, especially at high concentrations, could induce thermal aggregation of duck egg white protein, which could determine the characteristics of salted egg white after heating.  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