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.
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   

2.
 Cactus pear (Opuntia ficus-indica) juice with a total soluble solids content of 15.94 °Brix was fermented using the wine yeast Saccharomyces cerevisiae. A 94.54% conversion of fermentable sugar was achieved with an ethanol production of 55.3 ml/l. The pigment degradation was found to be 17% at the end of the time allowed for fermentation. However, the fermentation had actually ceased due to depletion of fermentable sugars after 12 h, a point at which only 9.4% pigment degradation was observed and there was no further total soluble solids degradation. The thermal stability of the yellow-orange pigment of the fermented juice was determined as a function of temperature at pH 5.0. The kinetic experiments were carried out at three different temperatures, 50, 70 and 90  °C. For a pseudo-first order thermal degradation rate the reaction rate constants were determined to be 0.0066, 0.0206 and 0.1244 min–1 for temperatures of 50, 75 and 90  °C, respectively. The activation energy was calculated as 15.71 kcal mole–1. The fermentation process did not affect the thermostability of the pigment extract. Received: 10 January 2000 / Revised version: 31 March 2000  相似文献   

3.
The diffusion of acetic or propionic acids from thin (44 to 54 μm) chitosan‐based antimicrobial packaging films in which they were incorporated was measured after immersion of the films in water, and the effects of pH (5.7, 6.4, or 7.0) and temperature (4 °C, 10 °C, or 24 °C) on diffusion were investigated. The kinetics of acetic‐ and propionic‐acid release deviated from the Fickian model of diffusion. Diffusion was found to be unaffected by pH in the range of values tested, but a decrease in temperature from 24 °C to 4 °C resulted in a reduction of diffusion coefficients from 2.59 × 10?12 m2.s?1 to 1.19 × 10?12 m2.s?1 for acetic acid and from 1.87 × 10?12 m2.s?1 to 0.91 × 10?12 m2.s?1 for propionic acid. The effect of temperature on diffusion was well (r2 > 0.9785) described by an Arrhenius‐type model with activation energies of 27.19 J.mole?1 (acetic) and 24.27 J.mole?1 (propionic). Incorporation of lauric acid or essential oils (cinnamaldehyde or eugenol) into the chitosan film at the time of preparation produced a subsequent reduction in the diffusion of acetic or propionic acid, and maximum effects were obtained with lauric acid and cinnamaldehyde incorporated to final concentrations of 1.0% and 0.5% (w/w), respectively.  相似文献   

4.
Thermal Inactivation Kinetics of Wheat Germ Lipoxygenase   总被引:2,自引:0,他引:2  
Thermal inactivation curves for wheat germ lipoxygenase (LPO) in partially purified and crude extracts were determined in capillary tubes at 60–68°C. The biphasic curves fitted a two-fraction first order model suggesting the presence of 2 groups of isozymes. At 60°C, the inactivation rate constants were 9.112 × 10?-5 set?1 and 9.174 × 10?-6 set?1 respectively, for thermolabile phase I and thermostable phase II in the partially purified extract, a difference of one order of magnitude. For the temperature change from 60 to 68°C, the rate constants increased by three orders of magnitude, implying a very high sensitivity (for LPO inactivation in partially purified extract ΔH?= 646261 J.mole?1, ΔS?= 1619 J.mole?1.K?1 for phase I, ΔH?= 546099 J.mole-l, ΔS?= 1298 J.mole?1.K-1 for phase II) to heat by both phases, although phase I was clearly the least stable.  相似文献   

5.
The kinetics of adsorption of patulin on activated carbon were studied at different initial patulin concentrations (100–400 ppb) for the temperature range 20–80°C. Apparent adsorption rate constants (kaapp) were changed from 1.07 × 10?3 to 1.86 × 10?3 g?1 min?1 while the temperature increased from 20 to 80°C. For equilibrium adsorption curves; the Langmuir model was attempted and model parameters (K and Q°) were obtained for different temperatures. Energy of activation and heat of adsorption were determined in a batch adsorption system (Ea= 2.02 kcal/mol and ΔH = 2.24 kcal/mol). The adsorption occurred endothermically and by physical mechanisms.  相似文献   

6.
Degradation of sorbic acid in aqueous glycerol solutions at pH 4·0 over the aw range 0·71–1·00 and the temperature range 40°–60°C was found to follow first-order reaction kinetics and to conform to the Arrhenius equation. Activation energy values obtained were 5·8 kcal mol?1 and 7·8 kcal mol?1 for systems at 0·80 aw with and without added Co++, respectively. The rate of sorbic acid degradation was observed to increase with decreasing aw (i.e. increasing glycerol concentration). The presence of added Co++ decreased the rate of sorbic acid breakdown at any particular aw or temperature. Browning of sorbate solutions during storage was markedly inhibited by Co++.  相似文献   

7.
 Thermal degradation of green asparagus ascorbic acid in high-temperature short-time conditions was studied by heating in a five-channel computer-controlled thermoresistometer. Ascorbic acid was heated to between 110  °C and 140  °C and the degradation kinetics were analyzed assuming that two different inactivation mechanisms were occurring, one aerobic and the other anaerobic. The two reactions followed first-order kinetics, with E a=12.3(2.0) kcal/mol and k 125  °C=47.0(3.0)×10–3 s–1 for the aerobic oxidation, and E a=6.1(1.4) kcal/mol and k 125  °C=4.1(0.2)×10–3 s–1 for the anaerobic degradation. Received: 30 January 1998 / Revised version: 11 June 1998  相似文献   

8.
Ascorbic acid (AA) degradation and colour changes, measured by the lightness index (L*), were determined in cashew apples (at low dissolved O2 concentrations) heated at high temperature (100–180 °C) in a hermetically sealed cell. A nonisothermal method was developed to estimate thermal degradation kinetics. The results showed that reaction kinetics during heat treatments were well represented by first‐order reactions. The temperature dependence of the kinetic constants was described by an Arrhenius type equation. The activation energy (Ea) for AA degradation and lightness index were 94 ± 3 and 98 ± 3 kJ mol?1, respectively. The reaction rate constant at 140 °C for AA degradation (64 × 10?5 ± 3 × 10?5 s?1) was twice that for the lightness index change (33 × 10?5 ± 2 × 10?5 s?1). Results allow generating temperature profiles of heat processes that would help preserve the AA of cashew apples as well as control the colour formation during high‐temperature processes.  相似文献   

9.
Kinetic parameters for thermal destruction of immunoglobulin A (IgA), lactoferrin, thiamin and folic acid in human milk were determined. Degradation proceeded following first order reaction kinetics. The times for 90% degradation (D value) at 60°C were, in seconds, 4.9 × 104 (68°C C 78°C) for IgA; 2.4 × 103 (58°C 70°C) for lactoferrin; 7.7 × 105 (95°C C 110°C) for thiamin and 1.9 × 104 (62°C C 78°C) for folic acid based on inactivation data at four constant temperatures between the range indicated. Z values (temperature change to alter degradation rate by a factor of 10) were 5.5°C, 4.7°C, 28.4°C, and 6.4°C for IgA, lactoferrin, thiamin and folic acid, respectively.  相似文献   

10.
The effects of soluble solids content and temperature on thermal properties of papaya puree were studied. Density and specific heat were measured using a pycnometer and differential scanning calorimeter, respectively, while thermal conductivity was measured using a line heat source probe. Thermal diffusivity was then calculated from the experimental results of the specific heat, thermal conductivity, and density. Thermal properties of papaya puree were experimentally determined within a soluble solids content range of 10 to 25 °Brix and temperature between 40 and 80°C. The density, specific heat, thermal conductivity, and thermal diffusivity of papaya puree were found to be in the ranges of 1014.6 to 1098.9 kg/m3, 3.652 to 4.092 kJ/kg °C, 0.452 to 0.685 W/m °C, and 1.127?×?10?7 to 1.650?×?10?7 m2/s, respectively. Moreover, the empirical models for each property as a function of soluble solids content and temperature were obtained.  相似文献   

11.
The fermentation of macerated tea leaf in aqueous suspension has been studied both manometrically and in a suitably controlled fermentation vessel. Oxygen consumption proceeded rapidly at a steady rate for an initial period, termed the primary fermentation, then decreased sharply as the availability of simple polyphenol substrates became rate limiting. Maximum reaction rates were observed at 20-35 mg ml?1 total solids, as a result of substrate inhibition. At higher concentrations of solids reaction rates were also affected by diffusion effects within the leaf particles. At 50 mg ml?1 total solids and 25°C the system approximated to Michaelis-Menton kinetics with Km02 = 5 × 10?5M and Vmax corresponding to a primary fermentation time (PFT) of 10-15 min. The PFT corresponded with the attainment of maximum total colour and maximum level of theaflavins (TF) (often >30% of total colour). Beyond the PFT total colour remained constant, TFs decreased and non-dialysable pigments increased with time. At fixed levels of dissolved oxygen and solids concentration > 60 mg ml?1 the proportion of TFs was inversely related to solids concentration. At 50 mg ml?1 total solids TFs increased to ca 60% of total colour as the dissolved oxygen level was increased to 0.175 mM. As temperature was increased from 15-35°C the reaction rate increased, the proportion of TFs decreased slightly and thearubigins of intermediate molecular weight increased markedly.  相似文献   

12.
The effect of freezing on ascorbic acid losses during water heating of potato cylinders at 50°, 65° and 85°C was studied at different periods of time. Loss of ascorbic acid in frozen potato was mainly by diffusion and greater than for fresh potato. The estimated apparent diffusivities of ascorbic acid in frozen potato during water blanching were 13.64 × 10-10 m2/sec at 50°C, 17.78 × 10-10 m2/sec at 65°C and 20.30 × 10-10 m2/sec at 80°C. The activation energy of the diffusion process was 3,007 cal/mol. Ascorbic acid aparent diffusion coefficients in potato tissue with pre-freezing treatment were between 80 and 90% of ascorbic acid diffusion coefficients in water in the 50–80°C range.  相似文献   

13.
The kinetics of ascorbic acid autoxidation was studied at pH 6.1 (μ= 0.08) using a buffered model system. Oxygen replenishment in the solution was accomplished by shaking the reaction vessel in a shaker water bath or by bubbling oxygen-nitrogen gas mixtures through the solution. Using the former system, a second order rate constant of 58.19 ± 11.92 min?1 M?1 was calculated (45°C), Since the shaker bath did not assure oxygen replenishment and dissolved oxygen levels could not be varied, the latter system was developed. The second order rate constants in the gas mixture experiments ranged from 20.80 ± 2.28 min?1 M?1 (30°C) to 176 ± 6.93 min?1 M?1 (55°C). Activation energies measured under different conditions ranged from 9.47 to 16.43 kcal mole?1.  相似文献   

14.
A tubular hyperfiltration membrane system was studied with tomato juice in a recirculation mode. Two aromatic polyamide thin film composite membranes (ZF 99) (1.3 m2) were examined. After 52 hr of operation with tomato juice one membrane exhibited a flux of 39.7 Lm?2hr?1 (72°C, 55.2 bar). Another membrane, after 717 hr, gave a flux of 41 Lm?2hr?1 (78°C, 41.4 bar) and natural tomato soluble solids (NTSS) rejection of 94 to >98%.  相似文献   

15.
Effects of temperature on the autoxidation and antioxidants changes of soybean, sunflower, and olive oils were studied. The oils were oxidized in the dark at 25, 40, 60, and 80 °C. The oil oxidation was determined by peroxide (POV) and p-anisidine values (PAV). Polyphenols and tocopherols in the oils were also monitored. The oxidation of oils increased with the oxidation time and temperature. Induction period decreased with the oxidation temperature; 87 and 3.6 days at 25 and 60 °C, respectively, for sunflower oil. The activation energies for the autoxidation of soybean, sunflower, and olive oils were 17.6, 19.0, and 12.5 kcal/mol, respectively. Olive oil contained polyphenols at 180.8 ppm, and tocopherols were present at 687, 290, and 104 ppm in soybean, sunflower, and olive oils, respectively. Antioxidants were degraded during the oil autoxidation and the degradation rates increased with the oxidation temperature of oils; for tocopherols, 2.1 × 10−3 and 8.9 × 10−2%/day at 25 and 60 °C, respectively, in soybean oil.  相似文献   

16.
The kinetics of the thermal degradation of vitamin C of marula, mango and guava pulp at different heat treatments at temperature ranging from 80 to 150 °C were investigated. For temperatures lower than 125 °C, the ascorbic acid in marula pulp was about 15 fold more stable to heat than the ascorbic acid in mango and guava pulp. The results showed that a simple first order degradation model could not describe the vitamin C degradation as biphasic behaviour was observed. Therefore the model was transformed in a two-fraction model in which the vitamin C content is divided in relatively stable and instable fractions. Marula had a low kd1,100°C of 7.2 × 10?3 min?1 compared to kd1,100°C of 1.2 × 10?1 min?1 for guava and 1.3 × 10?1 min?1 for mango. Guava had the highest activation energy, Ea of 58 kJ/mol, followed by mango with 39 kJ/mol and then marula with 29 kJ/mol.  相似文献   

17.
An amylase inhibitor isolated from black beans (Phaseolus vulgaris) can completely inhibit porcine pancreatic α-amylase forming a 1:1 stoichiometric complex. The kinetic pattern of complex formation is pH dependent. At pH 5.5 it follows a first order reaction with rate constant of 0.029 min?1 and 0.017 min?1 at 37°C and equimolar inhibitor and enzyme concentration, respectively, of 10?8 M and 10?9 M. At pH 6.9 it is a second order reaction, with a rate constant of 0.25 × 106 M?1 min?1 at 37°C, with 4 × 10?8 M concentrations of enzyme and inhibitor. The dissociation constants of the enzymeinhibitor complex are 1.7 × 10?10 M at pH 5.5 and 4.4 × 10?9 M at pH 6.9, at 37°C. The kinetic data obtained at pH 5.5 suggested the formation of an initial reversible complex followed by a conformational change step. The complex can be dissociated either in acid pH (4.3) or at pH values higher than 6, 5 with partial recovery of the amylase activity.  相似文献   

18.
It has been shown by a potassium permanganate titration method that solutions of egg white decompose hydrogen peroxide. Using an oxygen electrode the 1st-order rate constants for the decomposition of hydrogen peroxide (during the first minute of the reaction) by different samples of newly laid and laboratory heat-treated egg white, have been calculated. An Arrhenius plot of calculated denaturation constants has shown that the activation enthalpy, free energy and entropy changes required for the heat inactivation of the catalase-like property' were 39.8 kcal mole?1, 22.6 kcal mole?1 and 51 entropy units, respectively. The effect of heat on the lyspzyme, the a-mannosidase and the N-acetyl-β-D-glucosaminidase enzymes of egg white has also been studied and it has been shown that the activity of N-acetyl-β-D-glucosaminidase enzyme is reduced by heat at 53 to 57°. The activation enthalpy, free energy and entropy changes required for the heat inactivation of N-acetyl-β-D-glucosaminidase were 62.9 kcal mole?1, 21.8 kcal mole?1 and 124.5 entropy units, respectively. The results are discussed with particular reference to the occurrence, disputed by some workers, of a catalase enzyme in egg white, and to a possible application of the effects of heat on the egg white enzymes as a method for measuring the effectiveness of heat pasteurisation processes.  相似文献   

19.
A two-stage tubular hyperfiltration (reverse osmosis) pilot plant equipped with 15.6 m2 of aromatic polyamide thin film composite membrance (ZF 99) was studied with tomato juice (4–6% natural tomato soluble solids [NTSS]). It was operated over 100 hr in recirculation and single pass modes. The single pass mode was operated at33–38 bar (feed inlet pressure) and 70–93°C for 18.2 hr. Stage 1 flux was 48 to 34 Lm?2hr?1. Stage 2 flux was 30 to 20 Lm?2hr?1. Concentration went to 7.4–9.0% NTSS at 8.6 L/min, and to 6.3–6.4% NTSS at 18.6 L/min. NTSS rejection was >98%. Of permeate samples tested none showed detectable levels of sugars; only 25% showed detectable levels of organic acids.  相似文献   

20.
The effects of calcium (Ca2+) level in the soak water, blanch water and brine, blanching temperature, and total seed solids on dry bean canning quality were investigated to optimise a laboratory canning protocol. A linear increase in the Ca2+ level of soak water, blanch water and brine resulted in a linear decrease in hydration coefficient and percent washed drained weight but a linear increase in texture. Low Ca2+ level (10 mg kg−1) reduced the hydration time for dry bean seed from 14 to 1 h. Blanching temperatures of 50, 70 and 88 °C had non‐significant effects on canning quality traits. Blanching for 30 min at 70 °C for black bean or at 88 °C for navy bean and pinto bean resulted in percent washed drained weight ≥ 60, as required by the Canada Agricultural Products Standards Act. Seed solids levels of 95–97 g per 300 × 407 (14 fl oz) can were sufficient to attain a percent washed drained weight of 60. It was confirmed that the thermal processing conditions (115.6 °C retort temperature, 45 min) used in this study were sufficient to achieve commercial sterility. The optimised lab protocol for evaluation of the canning quality of dry bean breeding lines is as follows. Seed containing 95 g of solids for pinto bean, 96 g for navy bean and 97 g for black bean is soaked in water for 30 min at 20 °C and blanched for 30 min at 70 °C for black bean and 88 °C for navy bean and pinto bean in water containing 10 mg kg−1 of Ca2+. The seed is then transferred to a 300 × 407 can, filled with brine containing 10 mg kg−1 of Ca2+, 1.3% (w/v) of NaCl and 1.6% (w/v) of sugar. The can is then sealed, processed in steam at 115.6 °C for 45 min and cooled at 20 °C for 20 min. Cans are stored for at least 2 weeks prior to quality evaluation of the canned product. Canning of dry bean seed according to this protocol results in precise estimation of canning quality traits, particularly percent washed drained weight. © 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