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 15 毫秒
1.
The effect of potassium sorbate (0–2 g litre−1) and sodium nitrite (0–1 g litre−1) on the growth of four strains of Escherichia coli O157: H7 in tryptic soya broth at various pH levels (pH 4·0–7·0 for sorbate, pH 5·0–8·0 for nitrite) were determined at 37°C and 4°C. Among the pH levels tested, sorbate and nitrite exhibited the highest antimicrobial activity at pH 4·0 and 5·0, respectively. At pH 5·0 and 37°C, the presence of 500 mg litre−1 sorbate or 200 mg litre−1 nitrite completely inhibited the growth of E coli O157: H7. While at higher pH levels, 2 g litre−1 sorbate or 1 g litre−1, nitrite, the highest concentration tested, did not show significant antimicrobial action against the test organisms. At 4°C and pH 5·0, the inoculated test organisms did not showed any significant growth in preservative-free control media. Different degree of inactivation and injury was observed when E coli O157: H7 strain 933 was stored in TSB (pH 5·0) containing 1 g litre−1 sorbate or nitrite at 37°C. At 4°C, inactivation and injury of E coli O157: H7 cells was not observed in the medium containing sorbate or nitrite throughout the 24 h experimental period.  相似文献   

2.
Polyphenol oxidase (PPO) was extracted from Emir grapes grown in Turkey and its characteristics in terms of pH and temperature optima, thermal inactivation, kinetic parameters and potency of some PPO inhibitors were studied. The optimum pH and temperature for grape PPO were found to be 4.2 and 25 °C respectively using catechol as substrate. Km and Vmax values were found to be 25.1 ± 2.72 mmol L−1 and 0.925 ± 0.04 OD410 min−1 respectively. Of the inhibitors tested, the most potent was sodium metabisulfite, followed by ascorbic acid. The thermal inactivation curve was biphasic. Activation energy (Ea) and Z values were calculated as 251.4 kJ mol−1 (r2 = 0.996) and 8.92 °C (r2 = 0.993) respectively. Copyright © 2006 Society of Chemical Industry  相似文献   

3.
Glutaminase of Actinomucor taiwanensis was purified approximately 96-fold with a yield of 18%, by sequential fractionation with ammonium sul-phate, anion exchange with DEAE-Sepharose CL-6B and gel filtration with Sephacryl S-200. The pH and temperature optima of purified glutaminase were 8·0 and 45°C, respectively. Glutaminase was stable at a temperature up to 35°C and at pH values of 6·0–8·0. The molecular weight was 80000 as determined from SDS-PAGE. The enzyme activity was markedly inhibited by HgCl2. In the presence of 100 g litre−1 NaCl, the enzyme activity was inhibited 50%.  相似文献   

4.
The reversible unfolding reactions for phenylmethylsulphonyl fluoride (PMSF)-modified trypins from Atlantic cod (cod PMS-trypsin) and cattle (bovine PMS-trypsin) were monitored by fluorescence spectrophotometry as a function of urea concentration and temperature. For urea unfolding at 25°C, the free energy change at zero concentration of urea (ΔG(H2O)) for cod PMS-trypsin was 11(±4·4) kJ mol−1 compared with 18(±1·14) kJ mol−1 for bovine PMS-trypsin, while the mid-point concentration for urea unfolding curve ([urea]1/2) was 3·0(±0·57) M and 4·1(±0·16) M, respectively. From studies of enzyme heat unfolding, the mid point temperature of the thermal unfolding curve ( T m ) was 46(±1·4)°C for cod PMS-trypsin compared with 57(±2)°C for bovine PMS-trypsin. The standard free energy change (Δ ) for reversible thermal unfolding of cod PMS-trypsin was 9(±1) kJ mol−1 compared with 19(±1) kJ mol−1 for bovine PMS-trypsin. Values for the enthalpy (Δ H m ), entropy (Δ S m ) and heat capacity (Δ C p ) for heat unfolding are compared. Results from urea and thermal unfolding studies show that cod PMS-trypsin has a significantly lower conformational stability than bovine PMS-trypsin.  相似文献   

5.
ABSTRACT: Biodegradable poly(butylene adipate-co-terephthalate) (PBAT) films incorporated with nisin were prepared with concentrations of 0, 1000, 3000, and 5000 international units per cm2 (IU/cm2). All the films with nisin inhibited Listeria innocua, and generated inhibition zones with diameters ranging from 14 to 17 mm. The water vapor permeability and oxygen permeability after the addition of nisin ranged from 3.05 to 3.61 × 1011 g m m−2 s−1 Pa−1 and from 4.80 × 107 to 11.26 × 107 mL·m·m−2·d−1·Pa−1, respectively. The elongation at break (ɛb) was not altered by the incorporation of nisin (P > 0.05). Significant effect was found for the elastic modulus (E) and the tensile strength (σs) (P < 0.05). The glass transition and melting temperatures with the presence of nisin ranged from −36.3 to −36.6 °C and from 122.5 to 124.2 °C, respectively. The thermal transition parameters such as the crystallization and melting enthalpies and crystallization temperature were influenced significantly (P < 0.05) by incorporation of nisin into films. The X-ray diffraction patterns exhibited decreasing levels of intensity (counts) as the concentration of nisin increased in a range of 2θ from 8° to 35°. Formation of holes and pores was observed from the environmental scanning electron microscopy images in the films containing nisin, suggesting interaction between PBAT and nisin.  相似文献   

6.
《Food chemistry》2001,74(2):147-154
Polyphenoloxidase (PPO) of peppermint leaves ( Mentha piperita) was isolated by (NH4)2SO4 precipitation and dialysis. Its pH and temperature optima were 7.0 and 30°C, respectively. On heat-inactivation, half of the activity was lost after 6.5 and 1.5 min of treatment at 70 and 80°C, respectively. Sucrose, (NH4)2SO4, NaCl and KCl appeared to be protective agents of peppermint PPO against thermal denaturation. Km of this enzyme ranged from 6.25×10−3 M with catechol to 9.00×10−3 M with L-dopa. The I50 values of inhibitors studied on PPO were determined by means of activity percentage (I) diagrams. Values were 1.4×10−4 M, 1.7×10−4 M, 9.7×10−5 M, 2.45×10−4 M, 2.16×10−1 M, 1.83×10−5 M, 6.5×10−5 M, 1.4×10−2 M, 7.5×10−5 M, for potassium cyanide, glutathione, ascorbic acid, thiourea, sodium azide, sodium metabisulfite, dithioerythritol, β-mercaptoethanol and sodium diethyl dithiocarbamate respectively. Therefore, sodium metabisulfite was the most effective inhibitor.  相似文献   

7.
The heat stability of rapeseed 12S globulin (cruciferin) was examined using 8-anilinonaphthalene-1-sulphonic acid (ANS) as a fluorescence probe. Heating cruciferin (0·06–0·3 mg ml−1 in 10 mM glycyl–glycyl piperizine buffer, pH 7·0, with 0·1–1·0 M NaCl) for 20 min increased its hydrophobicity as monitored by ANS fluorescence measurements. The mid-point temperature for the heat effect (Tm) increased linearly with increasing solvent pH (Tm (°C)=4·16 pH+41 (μ=0.1)) or sodium chloride concentration (Tm (°C)=14·7 [NaCl]+71 (pH=7·0)). The range of Tm values for cruciferin was 45–96°C. At 20°C cruciferin was unstable at pH<3·0 but relatively stable under alkaline conditions (pH 8–10). Though possessing an oligomeric structure, cruciferin appears to heat denature in accordance with the two-stage deactivation model for simple globular proteins.  相似文献   

8.
The pullulan-hydrolyzing enzyme from the culture filtrates of Sclerotium rolfsii grown on soluble starch as a carbon source has been purified by ultrafiltration (Amicon, PM-10), ion-exchange chromatography (DEAE-Cellulose DE-52) and gel filtration chromatography (Bio-Gel P-150). The enzyme moved as a single band in non-denaturing polyacrylamide gel electrophoresis carried out at pH 2.9 and 7.5. The relative molecular mass of the enzyme was estimated to be 64.000 D by SDS-PAGE and 66.070 D by gel filtration on Bio-Gel P150. The enzyme hydrolyzed pullulan optimally at 50°C between pH 4.0–4.5, whereas, soluble starch was optimally hydrolyzed at a pH of between 4.0–4.5 and at 65°C. The Michaelis constant (Km) for pullulan was 5.13mg·ml−1 (Vmax 1.0U · mg−1) and for soluble starch, it was 0.6mg · ml−1 (Vmax 8.33 U · mg−1). The enzyme was observed to be a glycoprotein (12–13% carbohydrate by weight) and had a strong affinity for Concanavalin A. The enzyme hydrolyzed α-D-glucans in an exo-manner, which resulted in the release of glucose as the sole product of hydrolysis. Acarbose, a maltotetraose analog, was found to be a potent inhibitor of both pullulan and starch hydrolysis (100% inhibition at 0.06 μM). The enzyme has been characterized as a glucoamylase (1,4-α-D-glucan glucohydrolase, EC 3.2.1.3) showing a significant action on pullulan.  相似文献   

9.
《Food microbiology》1994,11(3):215-227
The effects of heating, thermoradiation and pH (5·5 to 7·0) on inactivation of V. vulnificus cells in buffers, oyster and fish homogenates were studied. Cells were more sensitive to thermoradiation than heating or radiation alone. Synergistic effects were observed during thermoradiation of V. vulnificus (107 cells ml-1) at 40°C in buffer (pH 5·5), in fresh oyster (pH 6·2) and in fresh fish (pH 6·7). This synergistic effect was also noted when the same number of cells in fresh fish homogenates were irradiated at 45°C. Inactivation of cells varied depending on the environment and were more pronounced in buffers than in oyster or fish homogenates. The D10 (dose in kGy inactivating 90% of cells) at 25·C was 0·078 in buffer (pH 7·0), 0·125 in oyster (pH 6·2) and 0·187 in fish (pH 6·7), but at 35°C, the D10 values were 0·054, 0·093 and 0·125 kGy, respectively. Low initial numbers of cells (10 ml-1) in pH 7·0 buffer were rapidly inactivated by thermoradiation (40·C) compared to high cell number (107 ml-1) and the D10 (kGy) was 0·024 for the former and 0·047 latter. These D10 values (kGy) were 0·046 and 0·093 in fresh oysters (pH 6·2), 0·093 and 0·109 kGy in fresh fish (pH 6·7), for the low and the high cell numbers, respectively following thermoradiation at 40°C.  相似文献   

10.
《Food microbiology》2001,18(3):299-308
The objective of this study was to determine the effect of warm, chlorinated water on the survival and subsequent growth of naturally occurring microorganisms and visual quality of fresh-cut iceberg lettuce. After dipping cut lettuce leaves in water containing 20 mg l−1free chlorine for 90 s at 50°C, samples were stored at 5 or 15°C for up to 18 or 7 days, respectively. Populations of aerobic mesophiles, psychrotrophs, Enterobacteriaceae, lactic acid bacteria, and yeasts and molds were determined. The visual appearance and development of brown discoloration were monitored. Treatment of lettuce in warm (50°C) chlorinated water delayed browning of lettuce. Shelf life of lettuce stored at 5°C, as determined by subjective evaluation of color and general appearance, was about 5 days longer than that of lettuce stored at 15°C. Treatment in warm (50°) water, with or without 20 mg l−1chlorine, and in chlorinated water at 20°C significantly (α= 0·05) reduced the initial population of mesophilic aerobic microflora by 1·73–1·96 log10cfu g−1. Populations increased, regardless of treatment, as storage time at 5°C and 15°C increased. The same trends were observed in populations of psychrotrophs and Enterobacteriaceae. Yeast populations increased slightly in lettuce stored at 5°C but were consistently about 3 logs lower than mesophilic aerobes. Populations of molds and lactic acid bacteria were less than 2 log10cfu g−1throughout storage at 5 or 15°C. Results suggest that heat (50°C) treatment may have delayed browning and reduced initial populations of some groups of micro-organisms naturally occurring on iceberg lettuce, but enhanced microbial growth during subsequent storage.  相似文献   

11.
Chinese cabbage (Brassica campestris L pekinensis group) was minimally processed using best preparation techniques and stored at 0 and at 5°C with and without dips in either citric acid, calcium chloride or ascorbic acid, all at 10 g litre−1. The visual quality, degree of chilling injury, pH and taste were evaluated. The most deleterious effects on quality were produced by black speck (gomasho) and browning. Citric acid inhibited the development of black speck and extended storage life from 10 days of the control to 14 days at 5°C. At 0°C the storage life was not extended by any dip, but citric acid improved quality by reducing black speck. Minimally processed Chinese cabbage treated with citric acid showed only a slight reduction of pH from 6·3 of the control to 6·1 (P⩽0·05) and taste was not significantly affected (P>0·05). Microbial spoilage was not apparent during storage at 0°C for 35 days and 5°C for 21 days under any treatment. © 1997 SCI.  相似文献   

12.
Nonenzymatic Browning in Pear Juice Concentrate at Elevated Temperatures   总被引:2,自引:0,他引:2  
The effect of temperature and soluble solids (°Brix) on nonenzymatic browning in pear juice concentrate was determined by following absorbance at 420 nm (A420) over the temperature range of 50–80°C. Browning could be modeled as a zero order rate process with rates of 22.2 × 10−4 (45.2 °Brix), 36.9 × 10−4 (55.4 °Brix), 53.5 × 10−4 (65.1 °Brix) and 107 × 10−4 (72.5 °Brix) A420· min−1 at 80°C. Temperature dependence was described by the Arrhenius relationship with an average activation energy of 21.9 kcal · mole−1. Formol titration indicated a 20% loss of amino acids during heating 4.4 hr at 80°C and no loss of carbohydrates was observed after any heating period.  相似文献   

13.
《Food chemistry》1999,64(3):351-359
Taro (C. esculenta) is a staple food in many tropical regions. A comparative study of crude polyphenoloxidases from taro (tPPO) and potatoes (pPPO) was carried out to provide information useful for guiding food processing operations. Crude PPO was prepared by cold acetone precipitation using ascorbic acid as antioxidant. The PPO content of taro acetone powder was 770±17 units (mg protein)−1 as compared with 3848±180 units (mg protein)−1 in potato acetone powder. The pH-activity optimum was pH 4.6 for tPPO and pH 6.8 for pPPO. Both enzymes retained >80% activity after incubation at pH 4.5–8 but there was rapid activity loss at pH < 4. The temperature-activity optimum (Topt) was 30°C for tPPO and 25°C for pPPO with 75 and 27% of their respective maximum activity retained at 60°C. Both tPPO and pPPO were irreversibly inactivated by 10 min heating at 70°C. The activation enthalpy (ΔH#) and activation entropy (ΔS#) for tPPO heat-inactivation were 87.4 (±0.1) kJ mol−1 and −56.2 (±4) J mol−1 K−1, respectively. For pPPO, ΔH# was 59.1 (±0.1) kJ mol−1 whilst ΔS# was −141 (±4) J mol−1 K−1. The apparent substrate specificity was established from values Vmax/Km as: 4-methylcatechol>chlorogenic acid>dl-dopa>catechol>pyrogallol> dopamine>>caffeic acid for tPPO. There was no detectable activity towards caffeic acid. The substrate specificity for pPPO was: 4-methylcatechol>caffeic acid>pyrogallol>catechol>chlorogenic acid >dl-dopa>dopamine. According to the order of inhibitor effectiveness (sodium metabisulphite>ascorbic acid>NaCl≈ (EDTA), there was a significant lag-phase before increases occurred in the absorbance at 420 nm. Preincubation of PPO with inhibitors increased the extent of inhibition, indicating a direct effect on the structure of the enzyme.  相似文献   

14.
The present study demonstrates the antiradical efficiency of myricetin, a flavonol widely distributed in fruits and vegetables, by testing its ability to react with two different free radicals, ABTS and DPPH·. The polyphenolic nature of myricetin led us to consider the possibility of its oxidation by polyphenol oxidase (PPO). The results reported show that myricetin can be oxidised by PPO extracted and partially purified from broad bean seeds. The reaction was followed by recording spectral changes with time, maximal spectral changes being observed at 372 nm. The presence of two isosbestic points (at 274 and 314 nm) suggested that only one absorbing product was formed. The spectral changes were not observed in the absence of PPO. The oxidation rate varied with the pH, reaching its highest value at pH 5.5. The myricetin oxidation rate increased in the presence of SDS, an activing agent of polyphenol oxidase. Maximal activity was obtained at 1.3 mM SDS. The kinetic parameters were also determined: V m = 1.35 µM min−1, K m = 0.3,mM , V m/ K m = 4.5 × 10−3 min−1. Flavonol oxidation was inhibited by a selective PPO inhibitor such as cinnamic acid (KI = 1 mM ). The results reported show that myricetin oxidation was strictly dependent on the presence of polyphenol oxidase. © 1999 Society of Chemical Industry  相似文献   

15.
ABSTRACT: The kinetics of anthocyanin degradation in blueberry juice during thermal treatment at 40, 50, 60, 70, and 80 °C were investigated in the present study. Anthocyanin degradation was analyzed up to the level of 50% retention using a pH differential method. The degradation of anthocyanin at each temperature level followed a first-order kinetic model, and the values of half-life time (t1/2) at temperatures of 40, 50, 60, 70, and 80 °C were found to be 180.5, 42.3, 25.3, 8.6, and 5.1 h, respectively. The activation energy value of the degradation of the 8.9 ° Brix blueberry juice during heating was 80.4 kJ·mol−1. The thermodynamic functions of activation (ΔG, ΔH, and ΔS) have been determined as central to understanding blueberry degradation.  相似文献   

16.
Raw mango ( Mangifera indica L) seed kernels were found to contain tannins (56·5 g kg−1 DM), cyanogenic glucosides (64 mg kg−1 DM), oxalates (42 mg kg−1 DM) and trypsin inhibitory activity (20 TIU g−1 DM). The contents of these anti-nutritive factors were lowered by both soaking and boiling treatments, but boiling was more effective. The in vitro protein digestibility (26·3%) and apparent metabolisable energy (7·88 MJ kg−1 DM) values of raw kernels were low, and these parameters were improved by soaking and boiling. The observed improvements paralleled reductions in tannin contents, indicating that tannins are largely responsible for the poor nutritive value of raw kernels. In experiment 1, diets containing 0, 50, 100, 150 and 200 g kg−1 raw mango seed kernels that replaced maize were fed to 7-day-old White Leghorn cockerels for 14 days. Inclusion of more than 50 g kg−1 raw kernels lowered ( P< 0·05) the weight gains, feed intake and feed efficiency of chicks. High level inclusion of raw kernels had toxic effects, as evidenced by increased mortality. In experiment 2, soaking and boiling treatments improved ( P< 0·05) feed intake of chicks fed on diets containing 100 g kg−1 kernels and reduced mortality. Feed/gain was unaffected by processing. Soaking had no effect, whereas boiling of kernels improved the weight gains. However, weight gains of chicks receiving diets containing 100 g kg−1 boiled kernels were numerically, though non-significantly, lower than those of the maize-control group. It is concluded that raw mango seed kernels are unsuitable as a feed ingredient in chick diets and, that soaking and boiling do not completely overcome the anti-nutritive effects of raw kernels.  相似文献   

17.
《Food microbiology》2000,17(3):269-275
Minimal thermal processing is desirable for near natural organoleptic and nutritional qualities of fruit based products. In the present investigation, the effect of heat (85°C) in combination with acidulants or common preservatives on inactivation of ascospores of Neosartorya fischeri, a heat resistant mould isolated from grapes, has been studied in mango and grape juice. The ascospores were found to survive for >300 min of heating at 70, 75 and 80°C in these fruit juices and complete inactivation required 120 min of heating at 85°C. The synergistic effect of heat and organic acids or preservatives in fruit juices was noticed. The thermal death rate (1/k85°C) values did not vary much in the presence of lactic (20), malic (20) and citric (19) acids, but tartaric acid showed least inactivation effect (1/k85°C=54 min) in mango juice. The 1/k85°Cvalues for ascospores of N. fischeri in mango juice containing 0·1% of potassium sorbate or sodium benzoate or combination of both at 0·05% were found to be 44, 35 and 29 min respectively. These values were respectively, 32, 13 and 14 min in grape juice. Nearly 50 and 67% of the heating time was reduced by the use of potassium sorbate and sodium benzoate (0·05% each) in mango and grape juice to inactivate 3 log number of ascospores of N. fischeri. These results may be useful in thermal processing of fruit juices.  相似文献   

18.
Freshly harvested beansprouts displayed a respiration rate of about 1 mmol O2 kg−1 h−1 at 10°C which was strongly dependent on temperature, a 10-fold increase being observed every 16·5°C (z=16·5°C, ie Q10=4·4). This commodity is also characterised by a high initial microbial load (about 107 cells g−1). During storage at various temperatures from 1 to 20°C, oxygen uptake rates dramatically increased with time and this phenomenon was well correlated with the development of aerobic microorganisms which reached 109 cells g−1 after 2 days at 20°C or 9 days at 1°C. Beansprouts were packaged in films, with permeabilities ranging from 950 to 200000 ml O2 m−2 day−1 atm−1, and stored at 8°C. Due to plant and microbial metabolism, oxygen concentrations decreased steadily within all packs until the onset of plant tissue decay. The latter occurred after 5–6 days with the least permeable films but did not occur within when the film permeability was over 100000 ml O2 m−2 day−1 atm−1. However, such films favoured brown discolouration, exudation texture and breakdown. The orientated polypropylene film (OPP) induced anoxic condition within 2 days and favoured anaerobic metabolism and necrosis of the sprouts. In all packages there was a rapid development of aerobic microorganisms and lactic acid bacteria that resulted in the accumulation of acetate and lactate and a decrease in pH. Thus, it clearly appeared that tissue decay was enhanced by microbial activity. At 8°C, 0·24 m2 of film per kg of sprouts provided the optimal atmosphere composition (ie 5% oxygen and 15% carbon dioxide) when a film permeability of 50000 ml O2 m−2 day−1 atm−1 was used. These conditions allowed a shelf-life of 4–5 days.  相似文献   

19.
Polyphenoloxidase (PPO, EC 1.14.18.1)was extracted from palmito (Euterpe edulis Mart) using 0.1 M phosphate buffer, pH 7.5. Partial purification of the enzyme was achieved by a combination of (NH4)2SO4precipitation (35–90% saturation) and Sephadex G-25 and DEAE-cellulose chromatography. The purified preparation gave five protein bands on polyacrylamide gel electrophoresis, three of them with PPO activity. The Kmvalues for chlorogenic acid, caffeic acid, catechol, 4-methylcatechol and catechin were 0.57, 0.59, 1.1, 2.0 and 6.25 mM , respectively. PPO has a molecular weight of 51 000 Da, maximum activity at pH 5.6 with chlorogenic acid as substrate, and was stable between pH 5.0 and 8.0. The enzyme was heat stable at 50–60°C and inactivated at 75°C. The heat stability of palmito PPO was found to be pH dependent; at 50°C and pH 4.0 the enzyme was fully inactivated after 30 min. The pH/activity studies showed two groups with pK values c 4.6 and 6.7 involved in PPO catalysis.  相似文献   

20.
The efficiency of crude extracelluar α-galactosidases from Cladosporium cladosporides, Aspergillus oryzae and A niger in reducing the raffinose and stachyose content in chickpea flours was studied and compared with other traditional treatments. The optimum pH for α-galactosidase activity was found to be 4·5 for A oryzae and 5·0 for Cl cladosporides and A niger, while the optimum temperature of enzyme activity was 40°C for Cl cladosporides and 50°C for A oryzae and A niger. The specific activities of α-galactosidase from Cl cladosporides, A oryzae and A niger were 3·35, 3·94 and 5·94 units μg−1 protein, respectively. The enzyme activity was stable between pH 4·0 and 7·0 for A oryzae and A niger and between pH 5·0 and 7·0 for Cl cladosporides. The enzymes were thermostable when incubated at temperature ranges of 40–60°C for Cl cladosporides and 40–50°C for A oryzae and A niger. The optimum conditions for removing the raffinose and stachyose were obtained by incubating chickpea flours with 30 ml of crude fungal α-galactosidase extract (290, 210 and 130 units ml−1 for Cl cladosporides, A oryzae and A niger, respectively) for 3 h at the optimum conditions of each strain. Crude fungal α-galactosidases reduced the raffinose oligosaccharides content in chickpea flours by 100%, while germination reduced the raffinose content by 69% and stachyose content by 75%. Other traditional techniques reduced the raffinose content by 13–49% and stachyose content by 10–32%. © 1998 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