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 31 毫秒
1.
In muscle food processing, where oxidation is inevitable, phosphates are usually added to improve water binding. This present study attempted to investigate the interactive roles of protein oxidation and pyrophosphate (PP) during thermal gelation of myosin. Myosin isolated from pork muscle was solubilized in 0.5 M NaCl at pH 6.2 then oxidatively stressed with an iron‐redox cycling system that produces hydroxyl radicals with or without 1 mM PP and 2 mM MgCl2 at 4 °C for 12 or 24 h then heated to 50 °C at 1.3 °C/min. Protein conformational stability was measured by differential scanning calorimetry, and covalent cross‐linking was examined by sodium dodecyl sulfate–polyacrylamide gel electrophoresis following chymotrypsin digestion. The binding of PP to myosin suppressed disulfide bond formation in myosin subfragments 1 and 2 and partially inhibited oxidation‐initiated cross‐linking of heavy meromyosin during myosin gelation with a lesser effect on light meromyosin. In the presence of PP, myosin exhibited less loss of conformational integrity upon oxidation than myosin without PP. Rheological analysis from 20 to 75 °C indicated up to 32% decreases (P < 0.05) in elastic modulus (G′) of myosin gels due to oxidation. However, the presence of 1 mM PP, which also lowered the gelling capacity of myosin, inhibited the oxidation‐induced G′ by nearly half (P < 0.05). These results suggest that the protection of myosin head from oxidative modification by PP can be a significant factor for the minimization of gelling property losses during cooking of comminuted meats.  相似文献   

2.
Antioxidant activity of soy protein (SP) and its hydrolyzed peptides has been widely reported. During scavenging of radicals, these antioxidative compounds would be oxidatively modified, but their fate is not understood. The objective of this study was to evaluate the structural characteristics of SP hydrolysates (SPHs), compared to intact SP, when used to neutralize hydroxyl radicals (?OH). SPHs with degree of hydrolysis (DH) 1 to 5 were prepared with Alcalase. Antioxidant activity of SPHs was confirmed by lipid oxidation inhibition measured with thiobarbituric acid‐reactive substances, ability to scavenge 2,2′‐azino‐bis(3‐ethylbenzothiazoline‐6‐sulphonic acid) radicals, and ferrous ion chelation capability. Oxidation of SPHs was initiated by reaction with ?OH generated from 0.1 mM FeCl3, 20 mM H2O2, and 1.0 mM ascorbate. After oxidative stress, carbonyl content of SPHs increased by 2‐ to 3‐fold and sulfhydryl groups decreased by up to 42% compared to nonoxidized samples (P < 0.05). Methionine, histidine, and lysine residues were significantly reduced as a result of inactivating ?OH (P < 0.05). Attenuated total reflectance‐Fourier transform infrared and circular dichroism spectroscopy suggested the conversion of helical structure to strands and turns. Oxidatively modified SPHs had a lower intrinsic fluorescence intensity but similar solubility when compared to nonoxidized samples. These structural changes due to ?OH stress may impact the ingredient interaction and functionality of SPHs in food products.  相似文献   

3.
BACKGROUND: The PC12 cell line is a suitable model for the investigation of neurodegenerative diseases. In this study, PC12 cells were used to examine in vitro antioxidative and antiapoptotic protection by S‐allyl cysteine (SAC), S‐ethyl cysteine (SEC) and S‐propyl cysteine (SPC). PC12 cells were treated with these agents at 5 and 10 µmol L?1 before exposure to hydrogen peroxide (H2O2). RESULTS: H2O2 treatment significantly decreased mitochondrial membrane potential (MMP) and cell viability and increased lactate dehydrogenase (LDH) release and DNA fragmentation (P < 0.05). The pre‐treatments with SAC, SEC and SPC significantly and concentration‐dependently elevated cell viability and MMP and lowered LDH release and DNA fragmentation (P < 0.05). H2O2 treatment also significantly increased levels of malondialdehyde (MDA), reactive oxygen species (ROS) and oxidised glutathione (GSSG) and decreased glutathione (GSH) content (P < 0.05). The pre‐treatments with SAC, SEC and SPC significantly decreased subsequent H2O2‐induced formation of MDA, ROS and GSSG (P < 0.05) and also alleviated H2O2‐induced GSH depletion (P < 0.05). Finally, H2O2 treatment significantly decreased Na+‐K+‐ATPase activity and elevated caspase‐3 activity (P < 0.05). The pre‐treatments with SAC, SEC and SPC significantly attenuated H2O2‐induced Na+‐K+‐ATPase activity reduction and caspase‐3 activity elevation (P < 0.05). CONCLUSION: The results obtained support that the three cysteine‐containing compounds studied are potent neuroprotective agents. Copyright © 2008 Society of Chemical Industry  相似文献   

4.
The objective of this study was to determine structure and functional properties changes of oxidised egg white protein derived from hydrogen peroxide (H2O2)/ferric chloride (FeCl3) /ascorbic acid hydroxyl radical‐generating systems at room temperature, including carbonyls, sulfhydryl and total sulfhydryl groups, dityrosine, free amino, surface hydrophobicity, particle size distribution, intermolecular forces, and foamability and emulsibility. Protein carbonyl and dityrosine content of egg white protein were increased (P < 0.05) with increasing the concentrations of H2O2. Oxidation decreased free sulfhydryl, total sulfhydryl groups and surface of hydrophobicity comparing with nonoxidised egg white protein. Oxidation also reduced its free amino content (P < 0.05). The low concentration of H2O2 contributing the average particle size of egg white protein was smaller than the control group. However, the high concentrations of H2O2 caused that egg white protein aggregated and the average particle size became larger. To sum up, oxidation made EWP denaturation and aggregation.  相似文献   

5.
The effects of oxidation on the chemical characteristics of myosin, β‐lactoglobulin and soy 7S globulin were investigated in a free radical‐generating system (FeCl3/H2O2/ascorbate). Oxidised myosin exhibited a marked increase (>10‐fold) in carbonyls, a small increase in amines and conversion of some free sulphhydryls to disulphide bonds. Oxidised β‐lactoglobulin and 7S globulin also showed a major increase in the carbonyl content (five‐ and threefold respectively), but other chemical changes were relatively minor. In the oxidised myosin/β‐lactoglobulin composites, some cross‐linked aggregates were formed. Aggregation was also evidenced in the myosin/7S globulin composites exposed to the oxidising agents. The results indicated that oxidation enhanced interactions of the non‐muscle proteins with myosin apparently through the modification of amino acid side chains. These physicochemical changes may influence the functionality of processed muscle foods. © 2000 Society of Chemical Industry  相似文献   

6.
Sweet potato starches were modified with three different concentrations of phosphorus oxychloride (POCl3) (0.01, 0.02, and 0.03%, based on dry weight of starch) as a cross‐linking agent. The effects of crosslinking on rheological and thermal properties of sweet potato starch (SPS) pastes were evaluated. Cross‐linking considerably reduced the swelling power, consistency index (K), apparent viscosity (ηa), and yield stress (σoc) values of SPS, which significantly decreased with increase in POCl3 concentration. The gelatinization temperature (Tp) and enthalpy (ΔH) values of the cross‐linked SPS, which were determined using differential scanning calorimetry, were higher than those of native SPS. Storage modulus (G′), loss modulus (G″), and complex viscosity (η*) of the cross‐linked SPS pastes determined using small deformation oscillatory rheometry, were higher than native starch, and they also decreased with increase in POCl3 concentration from 0.01 to 0.03%. The tan δ (ratio of G″/G′) values (0.15–0.19) of the cross‐linked SPS samples were much lower than that (0.37) of the native SPS, indicating that the elastic properties of the SPS pastes were strongly influenced by modifications from cross‐linking. Finally, Cox–Merz plots showed that η* was much higher than ηa for the cross‐linked SPS pastes.  相似文献   

7.
Protective effects and actions from s‐ethyl cysteine (SEC) and s‐methyl cysteine (SMC) for BEAS‐2B cells were examined. BEAS‐2B cells were pretreated with SEC or SMC at 4, 8, or 16 μmol/L, and followed by hydrogen peroxide (H2O2) treatment. Data showed that H2O2 enhanced Bax, caspase‐3 and caspase‐8 expression, and declined Bcl‐2 expression. However, SEC or SMC dose‐dependently decreased caspase‐3 expression and reserved Bcl‐2 expression. H2O2 increased reactive oxygen species (ROS) production, and lowered glutathione level, glutathione peroxide, and glutathione reductase activities in BEAS‐2B cells. SEC or SMC pretreatments reduced ROS generation, and maintained glutathione redox cycle in those cells. H2O2 upregulated the expression of both p47phox and gp91phox. SEC and SMC downregulated p47phox expression. SEC or SMC at 8 and 16 μmol/L decreased H2O2‐induced release of inflammatory cytokines. H2O2 stimulated the activation of nuclear factor‐κB (NF‐κB) and mitogen‐activated protein kinase. SEC and SMC pretreatments dose‐dependently downregulated NF‐κB p65 and p‐p38 expression. Pyrrolidine dithiocarbamate or SB203580 inhibited NF‐κB activation and p38 phosphorylation; thus, SEC or SMC pretreatments failed to affect protein expression of these factors. These novel findings suggest that SEC or SMC could protect bronchial cells and benefit respiratory epithelia stability and functions.  相似文献   

8.
Starches separated from different potato cultivars were modified using two different cross‐linking agents: epichlorohydrin (EPI) and phosphoryl chloride (POCl3) at different concentrations (1.0 and 2.0 g kg?1 POCl3; 2.5, 5.0 and 10 g kg?1 EPI). Differential scanning calorimetry, rheological and retrogradation measurements were performed to characterise the influence of cross‐linking on the properties of potato starches. Cross‐linking considerably reduced swelling power, solubility, water‐binding capacity and paste clarity. The decrease became greater as the reagent concentration increased. The starches treated with 1.0 g kg?1 POCl3 exhibited exceptionally higher swelling power than their counterpart native starches. Neither cross‐linking agent caused any change in morphology of the starch granules. Studies on the phase transitions associated with the gelatinisation showed significantly higher values for the onset temperature (To), peak temperature (Tp), conclusion temperature (Tc) and enthalpy of gelatinisation (ΔHgel) for the cross‐linked starches than the native starches. Starches treated with both the reagents showed lower peak storage modulus (G′) and loss modulus (G″) than their native counterparts. The tendency of the starch pastes towards retrogradation increased considerably with increases in storage duration. However, the starches treated with 1 g kg?1 POCl3 exhibited much lower syneresis than the other cross‐linked starches. Copyright © 2006 Society of Chemical Industry  相似文献   

9.
Cassava starch was cross‐linked with epichlorohydrin (EPI) at 45°C for 2 h in three different media which include water, water in the presence of a phase transfer catalyst (PTC) and N,N‐dimethylformamide (DMF). The products were characterized by determining their physicochemical, thermal and retrogradation properties. In aqueous medium, the use of a PTC, tetrabutylammonium bromide (TBAB) produced derivatives with higher degree of cross‐linking than those prepared without the use of the catalyst. The degree of cross‐linking was found to be higher using the same concentration of EPI when the reaction was carried out in DMF. At low levels of cross‐linking, the peak viscosity of the cross‐linked starches increased in comparison to that of the native starch. With increasing degree of cross‐linking, the peak viscosity showed a significant reduction. The swelling volume, solubility and light transmittance of the starch pastes were lower for the modified starches. The cross‐linked starches showed slightly reduced values for the gelatinization temperatures, Tonset, Tpeak and Tend. The enthalpy of gelatinization of the modified starches increased with increase in the degree of cross‐linking. The modified starches exhibited higher water‐binding capacities (WBC) than the native starch; but with increase in the degree of cross‐linking, there was a gradual decrease in WBC. The in vitro alpha amylase digestibility of the modified starches decreased gradually with increase in the level of cross‐linking.  相似文献   

10.
Cationic acetylated starch‐g‐poly(styrene‐butyl acrylate) surfactant‐free emulsion (CAS‐g‐poly(St‐BA)) was synthesized by graft copolymerization of styrene (St) and butyl acrylate (BA) onto CAS using FeSO4–H2O2 redox initiator. The maximum graft of 55.68% was derived when H2O2 concentration, monomer concentration, and St/BA ratio were 9%, 130%, and 1:1, respectively. The results obtained from FTIR, NMR (H1 NMR and C13 NMR), XRD, SEM, and thermogravimetric analysis (TGA‐DTG) confirmed graft copolymerization of St and BA onto CAS. And it was demonstrated that film‐forming properties of starch were greatly improved via grafting St and BA onto starch. It was also found that paper sized with CAS‐g‐poly(St‐BA) exhibited higher ring crush index and bursting strength than paper sized with cationic potato starch (CS) and CAS, as well as much lower water absorption, which is further verified by contact angles results.  相似文献   

11.
Lactic acid bacteria (LAB) have been used as ingredients of functional foods to promote health and prevent diseases because of their beneficial effects. This study aimed to investigate the antioxidative effects of LAB on the hepatotoxicity in D‐galactose‐induced aging mice. LAB were isolated from the traditional Chinese fermented foods and screened by the tolerance of hydrogen peroxide (H2O2). Male ICR (Institute of Cancer Research) mice were subcutaneously injected with D‐galactose for 5 weeks and then gastric gavage with LAB for 6 weeks. The results showed that Lactobacillus plantarum AR113 and AR501, and Pediococcu pentosaceus AR243 could tolerate up to 1.5 mM H2O2 in vitro, and they could live through simulated gastrointestinal tract (GIT) to colonizing the GIT of host. In vivo, oral administration of L. plantarum AR113 and AR501 improved the antioxidant status of D‐galactose‐induced oxidative stress mice such as alleviated liver damages and reduced abnormal activities of superoxide dismutase, glutathione peroxidase, and catalase to normal levels. In addition, L. plantarum AR501 markedly elevated the gene expression of nuclear factor erythroid‐2‐related factor 2 and upregulated the expressions of several antioxidant genes such as glutathione reductase, glutathione S‐transferase, glutamate‐cysteine ligase catalytic subunit, glutamate‐cysteine ligase modifier subunit, and NAD(P)H quinone oxidoreductase 1 in the liver of an aging mice. Therefore, L. plantarum AR501 could be a good candidate for producing antiaging functional foods.  相似文献   

12.
Hydrogels were synthesized through cross‐linking of carboxymethyl starch (CMS; Degree of Substitution DS = 0.45) using polyfunctional carboxylic acids (malic, tartaric, citric, malonic, succinic, glutaric and adipic acid). The syntheses used a cross‐linking agent ratio (ratio of the number of cross‐linking agent molecules to the number of monomer units constituting the polymer) of FZ = 0.05. After cross‐linking the gels were dried, ground and then hydrogels of a polymer concentration of 4 mass‐% were produced. These CMS‐hydrogels were then rheologically characterized using dynamic oscillatory measurements. From measurements of the plateau region storage modulus G'P, the network parameters molar mass between two entanglement points Me (Me ranging from 9.318 (citric acid) to 281.397 g/mol (tartaric acid)), the cross‐link density νe and the distance between two entanglement points ξ were calculated. Using carboxylic acids without other functional groups, a maximum in gel sturdiness is found at a spacer length of two CH2‐groups. The evaluation of the loss factor tan δ for the CMS‐hydrogels showed that values of tan δ = 0.2 varied only slightly with the frequency ω. Flow curves showed a pseudopIastic flow behavior for all CMS‐hydrogels (the shear viscosity η declining over five decades in the range of the shear rate γ of 10−3 to 103 s−1) The different polyfunctional carboxylic acids have a strong influence on the sturdiness of the synthesized CMS‐hydrogels.  相似文献   

13.
Treatment of mushroom tyrosinase with reducing agents such as hydrogen peroxide, ascorbic acid, phenylhydrazine, gallic acid, ferrocyanide and NH2OH resulted in inactivation of the enzyme. Under the conditions tested, 50% inactivation of the enzyme was obtained with 4 μM H2O2, 20 μM ascorbic acid, 40μM phenylhydrazine, 6 mM gallic acid, 12 mM ferrocyanide and 22 mM NH2OH. The ability of the reducing agents to reduce Cu2+ in a chemical model system was determined and it was found that gallic acid, phenylhydrazine, ascorbic acid and NH2OH are relatively good reductants of Cu2+ while H2O2 and ferrocyanide are relatively poor ones. The copper content of mushroom tyrosinase before and after inactivation by each of the reducing agents was determined. The copper content of the enzyme inactivated by H2O2, NH2OH, phenylhydrazine, ferrocyanide, gallic acid and ascorbic acid was 100%, 90%, 90%, 85%, 85% and 76% compared to that of the control (enzyme not treated). It was concluded that the degree of inactivation of mushroom tyrosinase by the reducing agents was not correlated with the decrease in the copper content of the enzyme nor with their ability to reduce Cu2+ in a chemical model system.  相似文献   

14.
Free‐radical scavenging capacities of strawberries and blackberries treated with methyl jasmonate (MJ), allyl isothiocyanate (AITC), essential oil of Melaleuca alternifolia (tea‐tree oil or TTO), and ethanol (EtOH) were investigated. All of these natural volatiles tested reduced the severity of decay in both strawberries and blackberries during storage at 10 °C as compared to the control. Most of these compounds enhanced antioxidant capacity and free‐radical scavenging capacity, except the AITC treatment. The MJ treatment for strawberries and blackberries had the highest antioxidant capacity, expressed as oxygen radical absorbance capacity (ORAC) values, after 7 days of storage. Moreover, the MJ treatment promoted the antioxidant capacity in strawberries and blackberries as measured by the radical 2,2‐di (4‐tert‐octylphenyl) ‐1‐picrylhydrazyl (DPPH) and the radical cation 2,2′‐azinobis(3‐ethylbenzothiazoline‐6‐sulfonic acid) diammonium salt (ABTS?+) scavenging activity in both 7 and 14 days after storage. The MJ treatment also increased scavenging capacities on the superoxide radical (O2??), hydrogen peroxide (H2O2), hydroxyl radical (?OH), and singlet oxygen (1O2) in strawberries and blackberries. Treatment with TTO or EtOH enhanced most of these free‐radical scavenging capacities, except for H2O2 in strawberries, and for O2?? and 1O2 in blackberries. These results indicated that all of the natural volatile compounds tested in this study, except AITC, promoted the antioxidant capacity and scavenging capacity of most major free radicals and, thus, helped to improve the physiology of berry fruits and enhanced their resistance to decay. While AITC was also very effective in reducing decay, its effect on free‐radical scavenging capacity was inconsistent, suggesting that additional mechanisms may be involved in its inhibition of fungal growth. Copyright © 2007 Society of Chemical Industry  相似文献   

15.
Zinc‐loaded whey protein nanoparticles were prepared by enzymatic cross‐linking whey protein followed by ethanol desolvation. Whey protein isolate (WPI) was denatured by heating (80 °C for 15 min) at pH 7.0 and then cross‐linked by transglutaminase at 50 °C for 4 h while stirring. Transglutaminase was inactivated by heating at 90 °C for 10 min, and then, ZnSO4·7H2O (1–10 mm ) was added. Zinc‐loaded whey protein nanoparticles were formed by adding ethanol at one to five times the volume of the protein solution at pH 9.0. The desolvated solutions were diluted by adding distilled water at ratio of 1:100 (w/v) immediately after desolvation. Dynamic light scattering (DLS) data showed that the particle size of zinc‐loaded whey protein nanoparticles increased with the amount of zinc and the volume of ethanol. Scanning electron microscopy micrographs revealed an almost spherical morphology for zinc‐loaded whey protein nanoparticles. The zinc loading efficiency was determined ranging from 76.7% to 99.2%. In vitro test data showed that the zinc release rate was low in simulated gastric fluid but high in simulated intestinal fluid. The results indicated that enzymatic cross‐linked whey protein nanoparticles may be used as a good vehicle to deliver zinc as a supplement.  相似文献   

16.
Oxidative effects on myosin, β‐lactoglobulin and soy 7S globulin as well as their composites were studied by differential scanning calorimetry and dynamic rheological analysis. Myosin, oxidised with a free radical‐generating system (FeCl3/H2O2/ascorbate), was destabilised, showing a marked decrease (>50%) in storage modulus during thermal gelation. On the other hand, oxidised β‐lactoglobulin and 7S globulin exhibited only minor changes in their thermal stability and gelation properties. In the non‐oxidised myosin/β‐lactoglobulin composite, β‐lactoglobulin affected the thermal denaturation and gelation properties of myosin through non‐covalent interactions. The nature and extent of the interactions were altered by oxidation. 7S globulin did not change the thermal stability but slightly enhanced the rheological properties of myosin, regardless of oxidative conditions. Hence, in muscle foods that contain whey or soy protein ingredients, all the protein components can be oxidatively modified, and the extent of the functionality changes in myosin will be influenced by the presence of the non‐muscle proteins. © 2000 Society of Chemical Industry  相似文献   

17.
The effects of prior enzymatic cross‐linking of bovine gelatin via horseradish peroxidase, glucose oxidase and glucose on microstructure and properties of the target film (cross‐linked gelatin film) were assessed. The cross‐linked gelatin film exhibited similar film thickness and moisture content, lower water solubility and higher opacity than the gelatin film directly prepared with bovine gelatin. The cross‐linked gelatin film also demonstrated improved thermal stability and mechanical properties, characterised by higher melting point and glass transition temperature, enhanced tensile strength and elongation at break and greater storage modulus. Prior gelatin cross‐linking resulted in 30.2% and 68.6% reductions in water vapour and CO2 permeability of the cross‐linked gelatin film, respectively, but did not affect oil permeability. Furthermore, the cross‐linked gelatin film possessed smaller cross‐sectional voids (diameter 100?360 nm vs. 200?595 nm) than the bovine gelatin film. This study shows that cross‐linking can efficiently improve film microstructure and properties of the gelatin‐like products.  相似文献   

18.
Myofibrils, oxidized with FeCl3/H2O2/ascorbate, exhibited an increase in carbonyls and amines, SH→SS conversion, peptide scission, myosin polymerization, and a decrease in thermal stability and gel‐formation ability. Amino‐acid side chains of whey‐protein isolates (WPI) and soy‐protein isolates (SPI) were also modified during oxidation, but the thermal stability of WPI or SPI was not significantly altered. Oxidation increased elasticity of SPI gel but not that of WPI gel. Similarly, oxidation promoted interactions of myofibrils with SPI but not with WPI, resulting in > 30% increases in elasticity of the myofibril/SPI composite gel over its nonoxidized control. Hence, in processed meats where oxidation occurs, the presence of soy proteins may enhance the functionality of myofibrillar proteins.  相似文献   

19.
Corn starch (20%, w/w) was non‐thermally and conventionally cross‐linked with phosphorus oxychloride (POCl3; 0.01, 0.05, or 0.1%, based on dry weight of starch) at 400 MPa for 5, 15 and 30 min and at 45°C for 2 h, respectively. Swelling power and solubility of both non‐thermally and conventionally cross‐linked corn starches were relatively lower than those of native corn starch. The pressure holding time did not affect the solubility and swelling power of non‐thermally cross‐linked corn starches. X‐ray diffraction patterns and relative crystallinity were not significantly altered by both conventional and non‐thermal cross‐linking. DSC thermal characteristics of both non‐thermally and conventionally cross‐linked corn starches were not significantly changed indicating that the double helical structure of amylopectin was not influenced by both conventional and non‐thermal cross‐linking reactions. Both non‐thermal and conventional cross‐linking greatly affected the Rapid Visco Analyser (RVA) pasting properties, such as increase in pasting temperature and decrease in peak viscosity compared to native starch. This result suggests that in case of cross‐linking using POCl3, both non‐thermal and conventional methods result in similar physicochemical properties and non‐thermal cross‐linking with POCl3 can reduce the reaction time from 2 h to 15 min. This work shows the potential and possibility of non‐thermal starch modification and provides the basic and scientific information on the physicochemical properties of non‐thermally cross‐linked corn starches with phosphorus oxychloride using UHP.  相似文献   

20.
Myofibrils isolated from post‐rigor chicken Pectoralis major (PM, white) and Gastrocnemius (Gas, red) muscles were irrigated with various concentrations of NaCl (0.1–1.0 M ) with or without 10 mM sodium pyrophosphate at pH 5.5 and 6.0. Structural changes were examined using phase contrast microscopy. PM myofibril samples tended to show more definitive H‐zones but obscure Z‐lines compared to Gas myofibrils. Significant myofibril swelling, accompanied by a pronounced protein extraction, occurred in 0.5 M NaCl solution. The extent of swelling as well as protein extraction increased with the NaCl concentration up to about 0.8 M . Addition of pyrophosphate facilitated myofibril swelling and reduced the minimal NaCl concentration for swelling to 0.4 M . Without pyrophosphate, protein extraction for both PM and Gas myofibrils occurred along the A‐band, sometimes starting from the centre, but when pyrophosphate was added, the extraction began from the ends of the A‐band. At pH 5.5, protein extraction was similar for PM and Gas, but at pH 6.0, PM myofibrils were more extractable and their architecture changed more extensively than Gas myofibrils, especially when pyrophosphate was present. The results may explain the different water‐imbibing abilities of white and red meat when processed with salt and phosphate. © 2000 Society of Chemical Industry  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