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.
The effect of limited hydrolysis was investigated on the physico‐chemical properties of soy protein isolate–maltodextrin (SPI‐Md) conjugate. The hydrolysates at a degree of hydrolysis (DH) of 1.8% showed much higher surface hydrophobicity (H0; 71.39 ± 3.60) than that of the SPI control (42.09 ± 2.17) and SPI‐Md conjugates (53.46 ± 2.74). Intrinsic fluorescence analysis demonstrated the unfolding of protein molecule and exposure of hydrophobic groups of SPI‐Md conjugate hydrolysates. As evidenced by far‐UV circular dichroism (CD) spectroscopy, the limited hydrolysis increased the unordered secondary structures of SPI‐Md conjugates. The denaturation temperature (Td) of SPI‐Md conjugate was significantly increased by subsequent limited hydrolysis from 102.53 ± 0.60 °C to 108.11 ± 0.61 °C at DH 1.8%. In particular, the emulsifying activity index (EAI) was improved notably after limited hydrolysis of DH 1.8% (147.76 ± 4.39 m2 g?1) compared with that of native SPI (88.90 ± 1.44 m2 g?1) and SPI‐Md conjugate (108.97 ± 1.45 m2 g?1).  相似文献   

2.
Milk protein concentrate was pretreated either by microwave irradiation or by ultrasound before initiation of 3‐h enzymatic hydrolysis. The duration of pretreatment ranged from 1 to 8 min at a power level of 800 W, with the control not being subjected to any pretreatment, and five enzymes (Alcalase, Trypsin, Neutrase, Alkaline Protease and Flavourzyme) were employed. The effects of microwave and ultrasound pretreatments on the kinetics and degree of hydrolysis, protein solubility, bitterness and angiotensin‐converting enzyme inhibitory activity were evaluated. Pretreatments increased the degree of hydrolysis and stabilised the solubility of the hydrolysates but could not significantly reduce bitterness of the hydrolysates The angiotensin‐converting enzyme inhibitory activity of the hydrolysates were improved with 5‐min ultrasound‐pretreated Neutrase hydrolysates giving IC50 value of 0.23 mg mL?1. Kinetic parameters showed improved catalytic efficiencies. Pretreatments of milk protein concentrates with either microwave or ultrasound significantly improve the bioactivity and functional characteristics of the resulting hydrolysates.  相似文献   

3.
Angiotensin I‐converting enzyme (ACE) inhibitory peptides from the shrimp shell waste (SSW) were isolated using different proteases. The orthogonal test results showed alcalase hydrolysates with ACE inhibitory activity of 67.07% under the optimal hydrolysis conditions of 60 °C hydrolysis temperature, pH = 9.5, 25 g L?1 substrate and 4000 U g?1 of enzyme, whereas neutral protease hydrolysates had an ACE inhibitory activity of 84.04% under the hydrolysis temperature of 50 °C at pH = 7.0 with 25 g L?1 of substrate and in the presence of 2000 U g?1 of enzyme. Neutral protease was more suitable for the production of ACE inhibitory peptides from SSW, where peptides with MW <5 kDa were recommended. The results of this study indicated that peptides obtained from SSW are as beneficial as antihypertension compounds in the functional food resources.  相似文献   

4.
Glutaminase (EC 3.5.1.2) was applied in this work to induce deamidation and hydrolysis of casein. Some reaction conditions based on casein deamidation were studied. Three casein hydrolysates with degree of deamidation of 2.8%, 5.8% and 8.5%, or degree of hydrolysis of 2.5%, 3.4% and 4.9%, respectively, were prepared at casein concentration 5% (w/v), glutaminase addition level 400 U kg?1 casein, reaction temperature 37 °C and reaction times 6, 12 and 24 h, respectively. Evaluation results showed that when iron (II) was added at 60 μm , iron (II)‐chelating powers of three hydrolysates were 41.1, 45.4 and 55.3%, while that of original casein and EDTA were 36.1 and 13.6%. Calcium (II)‐chelating power of three hydrolysates was 1.23, 1.41 and 1.49 mmol g?1 casein, whereas that of original casein was 1.05 mmol g?1 casein. Three hydrolysates also had ACE‐inhibitory activity in vitro, with IC50 values from 0.75 to 2.34 mg mL?1.  相似文献   

5.
Tea dregs possess abundant proteins, and the objective of this study was to investigate the antioxidant activity of tea dregs protein hydrolysate with limited hydrolysis by protamex and its possible action mechanism. Tea dregs protein was hydrolysed by alcalase, protamex or neutrase. The hydrolysis condition was optimised, and the hydrolysate was characterised for 1,1‐diphenyl‐2‐picryl hydrazyl (DPPH) radical‐scavenging activity, hydroxyl radical‐scavenging activity and antioxidant activity in linoleic acid (LA) system and in chicken products. Tea dregs protein hydrolysate (TDPH) was formulated (0.1%, 0.5%, 1.0%, w/w) into chicken products to determine in situ antioxidant efficacy. Thiobarbituric acid‐reactive substances (TBARS) and peroxide value (POV) formed in chicken products during storage (4 °C, 0–7 days) were analysed. Results showed that the optimum hydrolysis condition was at 50 °C, pH 7.0 for 20 min, and the concentration of tea dregs protein was 1.5%; ratio of protamex to substrate was 6000 U g?1. The radical‐scavenging ratio of TDPH to 1,1‐diphenyl‐2‐picryl hydrazyl (DPPH) was 90.30% at the concentration of 0.1 mg mL?1 and that to hydroxyl radical was 65.18% at the concentration of 1.0 mg mL?1. Moreover, it also showed strong antioxidant activity both in linoleic acid (LA) system and in chicken products. The molecular weight distribution of tea dregs hydrolysates was determined by nanofiltration tubular membrane, and the protein hydrolysates with molecular weight above 8000 Da had more effective antioxidant activity. The radical‐scavenging activities to DPPH and hydroxyl radical were 85.72% at 0.1 mg mL?1 and 71.52% at 1.0 mg mL?1, respectively. These findings suggest that the enzymatic hydrolysate of tea dregs protein probably possesses the specific peptides/amino acids which could stabilise or terminate the radicals through donating hydrogen. In addition, the hydrolysate could form a physical barrier around the fat droplets.  相似文献   

6.
Soya protein isolate (SPI) with or without added substrate‐like amino acid was subject to enzymatic hydrolysis catalysed by commercial proteases (Alcalase 2.4 L, flavourzyme and pancreatin). Addition of a small amount of amino acids (amino acid: SPI = 1: 2500, mol g?1) during hydrolysis would cause a significantly (< 0.05) reduced protein recovery, increased degree of hydrolysis, and altered amino acid composition and antioxidant activities of SPI hydrolysates. The SPI hydrolysates prepared with added Asp, Arg or Lys exhibited a higher antioxidant activity than the control. The bitterness of SPI hydrolysates was largely reduced upon addition of Met, Asp or Glu during hydrolysis, whilst the umami taste and mouthfeel‐liking were remarkably increased. Therefore, adding amino acid during hydrolysis is a feasible and beneficial approach to improve both the functional and sensory properties of SPI hydrolysate.  相似文献   

7.
The objective of the study was to determine the influence of potato variety, weight and storage time after lifting on the glycaemic index (GI) and resistant starch (RS) content predicted from measurement of the rate and extent of in vitro starch hydrolysis, respectively. The potatoes were either boiled, or boiled and subjected to different heat‐cycling conditions selected to promote retrogradation of amylose or amylopectin, respectively. The hydrolysis indices (HI) and predicted GIs of all 19 potato products were high and fell within narrow ranges of 122–144 and 118–138, respectively. No correlation between average weight of the potato tuber and HI was found. Furthermore, there was no difference in HI between potatoes stored for 1–3 or 8–10 months, nor between varieties of new potato and winter potato. However, the HI was significantly lowered by temperature cycling at conditions known to promote retrogradation of amylopectin (6 °C, 48 h) compared with 6 °C for 24 h followed by 70 °C for 24 h. RS content was already substantial in boiled potatoes, 4.5 g 100 g?1 (starch basis), and could be increased further by temperature cycling, the highest yield obtained, 9.8 g 100 g?1 (starch basis), following heat treatment at 6 °C for 24 h followed by 70 °C for 24 h; that is at conditions known to favour amylose retrogradation. Copyright © 2004 Society of Chemical Industry  相似文献   

8.
Effect of pretreatments with 1 w/v% sodium hydrogen sulphite (NaHSO3) and 1 w/v% calcium chloride (CaCl2) and drying temperatures (55, 60 and 65 °C) on sweet potato flour were investigated. Flour treated with CaCl2 had higher amounts of ascorbic acid and β‐carotene (10.61–12.54 and 3.26–3.46 mg 100 g?1 wet basis, respectively) than that treated with NaHSO3 (9.47–11.47 and 3.05–3.43 mg 100 g?1 wet basis, respectively). Total phenolic content and water absorption index (wet basis) were highest at 65 °C when treated with NaHSO3 (10.44 mg 100 g?1 and 2.49 g g?1 respectively) and CaCl2 (9.52 mg 100 g?1 and 2.85 g g?1 respectively). Swelling capacity (wet basis) was highest at 60 °C when treated with CaCl2 (2.96 g g?1) whereas when treated with NaHSO3 (2.85 g g?1) it was highest at 55 °C. Freeze‐dried samples treated with NaHSO3 had higher lightness and total phenolic content while CaCl2‐treated samples had higher β‐carotene and ascorbic acid. The results showed that good quality flour could be produced after soaking in CaCl2 and dried at 65 °C.  相似文献   

9.
Juice from mature-green ambarella contains starch, a characteristic detrimental to its visual appearance due to the white sediment formed upon storage. The purpose of this work was to evaluate the effects of starch and cell wall degrading enzymes on juice residual starch and content in soluble sugars. Starch and cell walls from mature-green ambarella fruits were purified and characterized. Starch was found to contain 21.0% amylose, 78.1% amylopectin and 0.9% other minors compounds. Its average granule size was 20 μm. Its thermal characteristics were: temperatures of onset (T o = 57.8 °C), peak (T p = 65.6 °C), and conclusion (T c = 72.6 °C) of gelatinization and the endothermic enthalpy (ΔH gelatinisation = 12.4 J g?1). Cell walls represented 2.8% of the edible fresh matter and were mainly constituted of highly methylated (HM) pectic substances and cellulose. The amylolytic preparations we tested, AMG® 300 L and Hazyme® C, showed similar behaviours in terms of starch hydrolysis and profit of Brix degree obtained. With 200 μg g?1 of AMG® 300 L or Hazyme® C, the degree of amylolysis of coarse ambarella puree was close to 50% and its increased up to 70% with enzymes concentrations higher than 1,000 μg g?1 (gelatinization at 75 ± 5 °C for 15 min followed by starch amylolysis at 60 ± 5 °C for 15 min). Total hydrolysis of ambarella starch is possible when pectinolysis occurred before amylolysis treatment, probably because of the fluidification of the medium by the pectocellulolytic enzymes. Pectinex® Ultra SP-L was the most efficient preparation to degrade the ambarella pectins (~80% of cell wall uronides liberated from 120 mg g?1 of purified cell walls within 1 h at 30 °C, pH 2.7).  相似文献   

10.
Rice bran contains 120–200 g kg?1 protein in addition to a large amount of fat, carbohydrate, and phytic acid. Rice bran protein (RBP) fractions were refined by a two‐step preparation to eliminate residual carbohydrate. The first step involved the sequential extraction of defatted rice bran into RBP fractions using their distinct solubility to give 37 g kg?1 of albumin, 31 g kg?1 of globulin, 27 g kg?1 of glutelin, and 2 g kg?1 of prolamin. In the second step, carried out by dissolving in respective solvent and isoelectric precipitation, the protein content of each fraction increased from 69% to 97% for albumin, from 71% to 90% for globulin, from 74% to 83% for glutelin, and from 18% to 20% for prolamin. The low protein content in the prolamin fraction might be due to its low solubility in the protein assay. Emulsifying stability index and surface hydrophobicity increased in the second‐step preparation of albumin and globulin, but not of glutelin. Emulsifying properties of RBPs were lower than that of a soybean protein isolate. Denaturation temperatures and enthalpy values of denaturation for albumin, globulin, glutelin, and prolamin were 50.1 °C/1.2 J g?1, 79.0 °C/1.8 J g?1, 74.5 °C/3.0 J g?1, and 78.5 °C/8.1 J g?1, respectively. No significant differences in the denaturation temperatures and enthalpy values of denaturation of RBP fractions were obtained with these two‐step preparations (P < 0.05). Copyright © 2007 Society of Chemical Industry  相似文献   

11.
Traditional processing technologies for fish by‐products containing significant amounts of oils usually either give high amounts of oil or maximised solubilisation of proteins. Due to lower yields and insufficient quality, the proteins or the oil is considered as secondary products. The proposed concept combines a gentle thermal separation of oil followed by enzymatic hydrolysis of the remaining protein‐rich fraction. The first stage, thermal treatment (40 °C) of fresh salmon backbones, separated up to 85% of the oil from the raw material and gave high‐quality oil (PV = 0.2 ± 0.0 meq kg?1, 0.16 ± 0.05% free fatty acids). Separation of a significant part of the oil gave reduced mass flow into the enzymatic stage, which then requires less enzymes and reduced energy consumption. Among the tested enzymes: Trypsin, Corolase PP and Mixture of Papain and Bromelain gave the highest yield of fish protein hydrolysates (FPH), while use of Protamex and Corolase PP resulted in FPH with the best sensory properties leading to the lowest bitterness.  相似文献   

12.
The stability of vitamin A in Refined Bleached Deodorised Palm Olein (RBDPOL) was studied for 24 months. Vitamin A decreased with time, temperature and thermal treatment (frying/cooking). RBDPOL fortification was observed over several temperature ranges, using PET, nylon and HDPE commercial packaging materials. After 24 months, the following vitamin A contents of 39–43 IU g?1 (39–45%) at 16–20 °C; 35–40 IU g?1 (43–49%) at 24–29 °C; and 28–39 IU g?1 (45–73%) at 24–45 °C were detected at the respective temperature ranges. Results showed stability of vitamin A fortified RBDPOL vegetable oil was not stable over typical shelf life (12 months). Depletion of vitamin A accelerated when the RBDPOL vegetable oil was subjected to high temperature thermal treatment.  相似文献   

13.
The aim of this study was to develop an enzymatic hydrolysis process of protein co‐products for two major commercial fish species in Australia: Atlantic salmon (AS) and Yellowtail kingfish (YTK). The outcomes are to produce high protein recovery of fish protein hydrolysates within controlled molecular weight ranges that display enhanced physicochemical properties of oil binding and emulsification. Three enzymes (Flavourzyme, Neutrase and Alcalase) were applied to processing co‐products. Protein recovery and physicochemical properties were evaluated with increasing hydrolysis time from 30 min to 180 min and ratio of enzyme to substrate (E/S) from 0.5% to 3.0%. In order to achieve a product with optimum emulsifying capacity (50 ± 0.6 mg?1), an E/S ratio of 0.6–1.3% Flavourzyme was applied for 30–111 min with a protein recovery of 55%; in order to achieve a product with optimum oil‐binding capacity (8.3 ± 0.3 g oil g hydrolysates?1), an E/S ratio of 2.3–3.0% Flavourzyme was applied for 25–64 min with a protein recovery of 70%. YTK protein hydrolysates were further membrane‐fractionated into five fractions (>100 kDa, 50–100 kDa, 30–50 kDa, 10–30 kDa and <10 kDa), and of these, the 10–30 kDa exhibited the best properties of oil binding (19 ± 0.3 g oil g hydrolysates?1) and emulsification (57 ± 0.7 mg?1). These results demonstrate the importance of enzymatic hydrolysis of seafood co‐products into high‐value ingredients for food products and processing.  相似文献   

14.
The angiotensin I-converting enzyme (ACE) inhibitory activity and hypocholesterolemic effect of Achatina fulica snail foot muscle protein hydrolysates (SFMPH) and its hydrolysates were studied. The SFMPHs were prepared at a temperature of 121°C for 60 min. To obtain the enzymatic hydrolysates, the SFMPHs were further hydrolysed with three proteases (papain, trypsin, or alcalase). Among all the hydrolysates, alcalase hydrolysate showed the highest degree of hydrolysis and was dominated by a small molecular size fraction (189–686 Da). The SFMPH treated by alcalase was effective in disintegrating intact cholesterol micelles. Furthermore, alcalase hydrolysate with a hydrolysis time of 60 min showed a strong ACE inhibitory activity in vitro with an IC50 of 0.024 mg/mL. Therefore, alcalase hydrolysate may be a promising ingredient for the use in functional foods.  相似文献   

15.
Sufu, a fermented soybean curd, was prepared by ripening the salted tofu cubes in the Aspergillus oryzae‐fermented rice–soybean koji mash at 25, 35 or 45 °C for a period of 16 days. It showed that the 16‐day ripened sufu contains less total isoflavone content (629.29–739.68 μg g?1 dried matter) than the salted tofu before ripening (942.59 μg g?1 dried matter). Regardless of ripening temperature, ripening causes a major reduction in the content of β‐glucoside and malonylglucoside isoflavones along with a significant increase of aglycone isoflavone content. These changes were enhanced as the ripening period extended. Among the various treatments examined, sufu ripened at 45 °C showed the greatest increase in aglycone content coupled with the greatest decrease in malonylglucosides. The distribution of malonylglucosides decreased from an initial 40.32% in the tofu cube to 9.78% after 16 days of ripening at 45 °C. Meanwhile, the distribution of aglycone increased from 13.17% to 39.88%.  相似文献   

16.
Optimisation of enzymatic hydrolysis of β‐casein with cell envelope proteinase (CEP) from Lactobacillus acidophilus JQ‐1 to produce the angiotensin‐I‐converting enzyme (ACE) inhibitory peptides using response surface methodology (RSM). Under optimal conditions (enzyme‐to‐substrate ([E]/[S]) ratio (w/w) of 0.132 and pH of 8.00 at 38.8 °C), the ACE inhibitory activity of hydrolysates was 72.06% and the total peptides was 11.75 mg mL?1. Scanning electron microscopy (SEM) micrographs indicated that the tightness of the β‐casein surface structure was gradually weakened and small holes appeared after enzymatic treatment, while Fourier transform infrared spectroscopy (FTIR) spectra indicated remarkable changes in the chemical composition and macromolecular conformation of β‐casein after enzymatic hydrolysis. Differential scanning calorimetry (DSC) analysis indicated that the corresponding hydrolysates had higher thermal stability. The enzymatic hydrolysis also led to an increase in the free sulfhydryl content of β‐casein hydrolysates compared with raw β‐casein, which led to the increase in the antioxidant activity of β‐casein hydrolysates.  相似文献   

17.
Thermal properties (thermal conductivity, thermal diffusivity and heat capacity) of Cheddar cheese were measured as a function of cheese age and composition. The composition ranged from 30–60% moisture, 8–37% fat, and 22–36% protein (wet basis). The thermal conductivity and heat capacity ranged from 0.354–0.481 W/m °C and from 2.444–3.096 kJ/kg °C. Both thermal conductivity and heat capacity increased with moisture and protein content and decreased with fat content. The thermal diffusivity ranged from 1.07×10?7 ? 1.53 × 10?7 m2/s. There was no significant relationship between thermal diffusivity and moisture, fat and protein content of cheese. No statistically significant effect (at the 10% level) of age (0 to 28 wk) on thermal properties was observed. Models predicting thermal properties as a function of cheese composition were developed and their predictive ability was compared with that of empirical models available in the literature. In addition, several theoretical thermal conductivity models were evaluated for their usefulness with Cheddar cheese. Published thermal conductivity models cannot accurately predict (mean error was from 3.4 to 42%) the thermal conductivity of Cheddar cheese.  相似文献   

18.
Response surface methodology was applied to optimise the aflatoxin reduction in both naturally and artificially contaminated samples using dry oven. The effect of initial aflatoxin concentration (0–400 ng g?1), heating time (30–120 min) and temperature (90–150 °C) was evaluated. The maximum reduction of AFB1 (78.4%) and AFB2 (57.3%) of artificially contaminated samples with initial aflatoxin concentration of 237 and 68 ng g?1, and those of AFG1 (73.9%) and AFG2 (75.2%) with initial aflatoxin concentration of 215 and 75 ng g?1 was obtained at 150 °C. The maximum reduction of AFB1 (80.2%) and AFB2 (69.7%) of naturally contaminated samples with initial aflatoxin concentration of 174 and 25 ng g?1 was obtained at 150 °C and 130 °C, respectively.  相似文献   

19.
Impacts of plastein reaction on bitterness, physicochemical and antioxidant properties of salmon frame hydrolysate with the aid of various proteases (alcalase and papain) at different concentrations and varying reaction temperatures were investigated. Plastein was produced from hydrolysate by papain at 40°C, which had 30% degree of hydrolysis (30DHP). Rearrangement of peptides in hydrolysate was performed by 1% papain at 40°C for 10 h, yielding plastein namely ‘30DHP-P1’. It showed the lowest bitterness (P < 0.05) than other plasteins and hydrolysates. Surface hydrophobicity was not related well with bitterness. Therefore, the size of peptides also determines the bitterness. 30DHP-P1 had augmented solubility; however, its antioxidant activities (DPPH and ABTS radical scavenging activities and ferric reducing antioxidant power) were slightly lower (P < 0.05) than those of hydrolysates. Bitterness of hydrolysate was markedly debittered via plastein reaction under optimal condition. Plastein generally had lighter colour and still possessed antioxidant activity.  相似文献   

20.
BACKGROUND: Enzymatic proteolysis of food proteins is used to produce peptide fractions with the potential to act as physiological modulators. Fractionation of these proteins by ultrafiltration results in fractions rich in small peptides with the potential to act as functional food ingredients. The present study investigated the angiotensin‐I converting enzyme (ACE‐I) inhibitory and antioxidant activities for hydrolysates produced by hydrolyzing Vigna unguiculata protein extract as well as ultrafiltered peptide fractions from these hydrolysates. RESULTS: Alcalase®, Flavourzyme® and pepsin–pancreatin were used to produce extensively hydrolyzed V. unguiculata protein extract. Degree of hydrolysis (DH) differed between the enzymatic systems and ranged from 35.7% to 58.8%. Fractionation increased in vitro biological activities in the peptide fractions, with IC50 (hydrolysate concentration in µg protein mL?1 required to produce 50% ACE inhibition) value ranges of 24.3–123 (Alcalase hydrolysate, AH), 0.04–170.6 (Flavourzyme hydrolysate; FH) and 44.7–112 (pepsin–pancreatin hydrolysate, PPH) µg mL?1, and TEAC (Trolox equivalent antioxidant coefficient) value ranges of 303.2–1457 (AH), 357.4–10 211 (FH) and 267.1–2830.4 (PPH) mmol L?1 mg?1 protein. CONCLUSION: The results indicate the possibility of obtaining bioactive peptides from V. unguiculata proteins by means of a controlled protein hydrolysis using Alcalase®, Flavourzyme® and pepsin–pancreatin. The V. unguiculata protein hydrolysates and their corresponding ultrafiltered peptide fractions might be utilized for physiologically functional foods with antihypertensive and antioxidant activities. Copyright © 2010 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