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 interaction of native calf thymus DNA (CT-DNA) with 2-tert-butyl-4-methylphenol (TBMP) at physiological pH has been investigated by spectrofluorometric and viscosimetric techniques. TBMP molecules were found to intercalate between base pairs of DNA, demonstrated by an increase in the specific viscosity of DNA and decrease in the fluorescence of TBMP solutions in the presence of increasing amounts of DNA and the calculated binding constants (K f) at different temperatures. Furthermore, the enthalpy and entropy of the reaction between TBMP and CT-DNA showed that the reaction is exothermic and enthalpy favored (ΔH = −19.18 kJ mol−1; ΔS = −26.98 J mol−1 K−1) which are other evidences to indicate that TBMP is able to be intercalated in the DNA base pairs.  相似文献   

2.
In the present study, the interaction of vanillin and human serum albumin (HSA) has been characterised by molecular modelling, fluorescence, Fourier transform infrared (FT-IR) and circular dichroism (CD) spectroscopic methods. The results of molecular modelling suggested that vanillin was located within the binding pocket of subdomain IIA of HSA mainly by hydrophobic forces. The quenching of HSA fluorescence takes place with a binding constant (K) of 8.8, 7.7, 5.7, 4.2 × 104 M−1 at four different temperatures (288, 298, 308, 318 K), respectively. Meanwhile, the number of binding site (n ≈ 1) was also obtained from fluorescence titration data. The enthalpy change ΔH0 and the entropy change ΔS0 were calculated to be −20 kJ mol−1 and 5.8 J mol−1 K−1 according to the Van’t Hoff equation. Furthermore, the alterations of protein secondary structure in the presence of vanillin were explored by FT-IR and CD spectra.  相似文献   

3.
Aqueous solubility of calcium citrate tetrahydrate was found to decrease with increasing temperature, while solubility of hexahydrate increased with a transition temperature at 51.6 °C. Excess citrate increased calcium citrate solubility but decreased the calcium ion activity of the saturated solution with an initial solubility overshooting to form supersaturated solutions indicating binding of calcium to citrate with an association constant of 3.6 ± 0.1 × 104, ΔHº = −5.07 ± 0.04 kJ mol−1, ΔSº = 70.3 ± 0.3 J mol−1 K−1 at 25 °C. Dissolution of the tetrahydrate and hexahydrate was found to have ΔHº = 27 ± 9 kJ mol−1, ΔSº = −218 ± 30 J mol−1 K−1 and ΔHº = 57 ± 7 kJ mol−1, ΔSº = −126 ± 24 J mol−1 K−1, respectively, as determined from the temperature dependence of solubility corrected for complex formation. The exothermic complex formation results in inverse solubility only for the tetrahydrate with its moderate endothermic dissolution, which also precipitates at ambient temperature rather than the less soluble hexahydrate.  相似文献   

4.
The interactions between the flavan‐3‐ol (?)‐epigallocatechin‐3‐gallate (EGCG) and bovine β‐casein in phosphate‐buffered saline (PBS) of pH 6.5 subjected to thermal processing at various temperatures (25–100 °C) were investigated using fluorescence quenching. The results indicated that different temperatures had different effects on the structural changes and EGCG‐binding ability of β‐casein. At temperatures below 60 °C, the β‐casein–EGCG interaction changed little (> 0.05) with increasing temperature. At temperatures above 80 °C, native assemblies of β‐casein in solution dissociated into individual β‐casein molecules and unfolded, as demonstrated by a red shift of the maximum fluorescence emission wavelength (λmax) of up to 8.8 nm. The highest quenching constant (Kq) and the number of binding sites (n) were 0.92 (±0.01) × 1013 m ?1 s?1 and 0.73 (±0.02) (100 °C), respectively. These results provide insight into the potential of interactions between β‐casein–EGCG that may modulate bioactivity or bioavailability to be altered during thermal process.  相似文献   

5.
The interaction of trypsin with acteoside was studied using ultraviolet visible absorption, fluorescence, synchronous fluorescence, circular dichroism techniques, along with molecular docking method. The fluorescence experiments indicated that acteoside quenched the intrinsic fluorescence of trypsin via a combined quenching process (static and dynamic quenching). The binding constant of acteoside to trypsin obtained was 2.50 × 105 L mol?1 at 298 K and the number of binding site was about one under the same experimental condition. The thermodynamic functions ΔH° and ΔS° of the binding process were 8.79 kJ mol?1 and 132.58 J mol?1 K?1, respectively, which indicated that the hydrophobic force was the main acting force between them. Ultraviolet–visible, synchronous fluorescence together with circular dichroism spectra studies demonstrated that the interaction of acteoside with trypsin lead to a loosening and unfolding of the protein backbone with partial β-sheet structures being transformed into α-helix structures. All these experimental results were validated and explained reasonably by docking studies. And the molecular docking results further illustrated that besides hydrophobic forces, hydrogen bonds also played an important role in the stabilization of the acteoside–trypsin complex. Results from this study should be helpful to make full use of acteoside in the food industry and be useful to the design of the drugs for the diseases related to trypsin.  相似文献   

6.
The standard static gravimetric method was used to determine moisture desorption isotherms (MDIs) of two onion varieties (Goudami and Galmi Violet) at 30 °C, 45 °C and 60 °C in the water activity ranging from 0.055 to 0.83. The combined effects of temperature and water activity on the antioxidant activities of the onion varieties were also studied. GAB, Oswin, Smith and BET equations were tested to fit the experimental data. The net isosteric heat of sorption was calculated. Equilibrium moisture content (EMC), total phenolic content (TPC) and antiradical activity were also measured. The isotherm and the EMC vary significantly with the onion variety and drying temperature, irrespective of water activity (aw). Desorption isotherms were best described by the GAB model. The maximum net isosteric heats for Galmi Violet (32.58 kJ mol?1) were greater than those of Goudami (23.50 kJ mol?1) at each EMC. The TPC and antiradical activity of the Galmi Violet were significantly (≤ 0.05) higher than that of the Goudami at all investigated temperatures and water activities.  相似文献   

7.
In this study, polyphenol oxidase (PPO) and vitamin C were used as the indicators of enzymes and nutrients to evaluate the apple quality during high humidity air impingement blanching (HHAIB) process. The PPO can be completely inactivated within 7 min at 90–120 °C and can retain relatively more vitamin C in the case of PPO fully inactivation. PPO inactivation followed zero‐order kinetics model at 90 and 100 °C, and followed first‐order fraction model at 110 and 120 °C. Activation energy (Ea) of PPO inactivation was between 11.61 and 13.66 kJ mol?1 by Arrhenius equation. Vitamin C degradation under all processing temperatures was well described by first‐order model and its Ea value was 26.69 kJ mol?1. Therefore, the HHAIB process was proved to be an effective pretreatment for Fuji apple quarters to inactivate PPO fast and meanwhile to maintain produce quality.  相似文献   

8.
Binding interactions of human serum albumin (HSA) with sunset yellow (SY) and allura red (AR), two food colorants, were investigated at the molecular level through fluorescence and UV absorption as well as molecular docking. The collective results of the study under the simulated physiological conditions proposed a static type of binding occurring between the two dyes and HSA. When compared with AR (293 K: Ksv = (4.21 ± 0.36) × 104 L·mol−1; Kb = (0.30 ± 0.23) × 106 L·mol−1), SY (293 K: Ksv = (6.80 ± 0.10) × 104 L·mol−1; Kb = (3.11 ± 2.01) × 106 L·mol−1) had stronger quenching ability and higher affinity for HSA due to less steric hindrance. It can be deduced that the energy transfer from HSA to the two dyes occurred with high probability based on the Förster resonance energy transfer theory (r < 7 nm, 0.5 R0 < r < 2.0 R0). The spectral analysis suggested that the formation of the dye-HSA complex resulted in the change in microenvironment around Tyr and Trp residues and in the secondary structure of the protein. According to molecular docking simulation, the two structural analogs almost bound to the same site of HSA, near Sudlow's Site I, but significant difference existed in the number and location of hydrogen bond (H-bond) formed between the dyes and HSA. From the molecular docking along with the thermodynamic parameters (AR: ΔHo = −(58.79 ± 15.24) kJ·mol−1, ΔSo = −(115.1 ± 31.10) J·mol−1·K−1; SY: ΔHo = −(52.24 ± 3.15) kJ·mol−1, ΔSo = −(50.07 ± 11.14) J·mol−1·K−1), it could be inferred that H-bond and van der Waals forces were the major binding forces involved in formation of the dye-HSA complexes.  相似文献   

9.
A dimeric serine protease Neriifolin S of molecular mass 94 kDa with milk clotting activity has been purified from the latex of Euphorbia neriifolia by anion exchange and size-exclusion chromatography. It hydrolyses peptidyl substrates l-Ala-pNA with highest affinity (Km of 0.195 mM) and physiological efficiency (Kcat/Km of 144.5 mM s). Enzyme belongs to the class of neutral proteases with pI value of 6.8, optimal proteolytic activity displayed at pH 9.5 and temperature 45 °C. Its proteolytic activity is strongly stimulated in the presence of Ca+2 ions and exclusively inhibited by serine protease inhibitors. Enzyme is fairly stable toward chemical denaturants, pH and temperature. The apparent Tm, was found to be 65 °C. Thermal inactivation follow first order kinetics with activation energy (Ea), activation enthalpy (ΔH∗), free energy change (ΔG∗) and entropy (ΔS∗) of 27.54 kJ mol−1, 24.89 kJ mol−1, −82.34 kJ mol−1 and 337.20 J mol−1 K−1.  相似文献   

10.
The process of extracting the astaxanthin oleoresin from pretreated Phaffia rhodozyma cells was optimised using a Box‐Behnken response surface design. Microwaving the cells at 105 W for 1 min followed by ethyl acetate extraction was the best pretreatment, and the optimal extraction conditions were 65 °C for 24 min using a solvent–solid ratio of 19:1. The order of the ability to disperse the astaxanthin oleoresin was propylene glycol> Tween 80 > Tween 20 > α‐cyclodextrin, β‐cyclodextrin. It was determined that the degradation of the colour of the water‐dispersible oleoresin followed a first‐order kinetics model. The greatest stability was observed at pH 4 and at the lowest temperature evaluated (40 °C). The thermal degradation of the pigment occurs in two steps, the first one from 0 to 1.5 h, with an Ea= 10.31 kJ mol?1, and the second one from 1.5 to 5 h, with an EaII = 30.06 kJ mol?1  相似文献   

11.
The interactions of digestive enzymes (pepsin, pancreatin) and milk proteins (β‐casein, β‐lactoglobulin (β‐Lg)) with (?)‐epigallocatechin gallate (EGCG), (?)‐epigallocatechin (EGC) and (?)‐epicatechin (EC) at gastric and intestinal pH were investigated by fluorescence spectroscopy. The results indicated that in the gastric environment, all three tea catechins showed binding affinities in descending order of strength with β‐casein first, followed by β‐Lg and then pepsin. The highest affinity was observed for EGCG–β‐casein, with a binding constant (KA) of 2.502(±0.201) × 105 m ?1. In the intestinal environment, the binding strengths of the proteins with EGCG and EGC were in the order β‐Lg > pancreatin > β‐casein; for binding with EC, the strength order was β‐casein > β‐Lg > pancreatin. The combination EGCG–β‐Lg had the strongest binding affinity, with a KA of 14.300(±0.997) × 105 m ?1. Thermodynamic analysis revealed that tea catechins complexed with milk proteins and digestive enzymes via different hydrophilic and hydrophobic interactions depending on the different digestion environments and types of catechins, proteins and enzymes.  相似文献   

12.
Mathematical modelling was used to study the effect of process temperature on moisture and salt mass transfer during osmotic dehydration (OD) of jumbo squid with 6% (w v −1) NaCl at 75, 85 and 95 °C. The diffusion coefficients for moisture and salt increased with temperature. Based on an Arrhenius-type equation, activation energy values of 62.45 kJ mol−1 and 52.14 kJ mol−1 for moisture and salt, respectively, were estimated. Simulations of mass transfer for both components were performed according to Newton, Henderson and Pabis, Page, Weibull and logarithmic mathematical expressions. The influence of drying temperature on the kinetic parameters was also studied. Based on statistical tests, the Weibull and logarithmic models were the most suitable to describe the mass transfer phenomena during OD of jumbo squid.  相似文献   

13.
Degradation kinetics of monomeric anthocyanins in acerola pulp during thermal treatment by ohmic and conventional heating was evaluated at different temperatures (75–90 °C). Anthocyanin degradation fitted a first-order reaction model and the rate constants ranged from 5.9 to 19.7 × 10−3 min−1. There were no significant differences between the rate constants of the ohmic and the conventional heating processes at all evaluated temperatures. D-Values ranged from 116.7 to 374.5 for ohmic heating and from 134.9 to 390.4 for conventional heating. Values of the free energy of inactivation were within the range of 100.19 and 101.35 kJ mol−1. The enthalpy of activation presented values between 71.79 and 71.94 kJ mol−1 and the entropy of activation ranged from −80.15 to −82.63 J mol−1 K−1. Both heating technologies showed activation energy of 74.8 kJ mol−1 and close values for all thermodynamic parameters, indicating similar mechanisms of degradation.  相似文献   

14.
Two extractions with methanol and water were used to determine the antioxidant and binding properties of some berries as a supplement to food. Fluorometry, FTIR spectra and radical scavenging assays were used for characterisation of bioactive compounds (polyphenols, flavonoids, flavanols and tannins) and the levels of their antioxidant activities (AAs). The contents of bioactive compounds and AAs in water and methanol polyphenol extracts in gooseberries, blueberries and cranberries differed, but not always significantly. Water extracts of gooseberries showed the lowest amounts of polyphenols (mg GAE g?1), 6.24 ± 0.6, and flavonoids (mg CE g?1), 0.29 ± 0.01, and AAs (μMTE g?1) determined by DPPH, FRAP, ABTS and CUPRAC assays such as 6.05 ± 0.6, 8.07 ± 0.9, 18.70 ± 1.8 and 13.44 ± 1.2, respectively, in comparison with blueberries and cranberries. Polyphenol content highly correlated with antioxidant activity (R2 from 0.94 to 0.81). The quenching properties of berries were studied by the interaction of water and methanol polyphenol extracts with HSA by 3D fluorescence. In conclusion, the bioactivity of gooseberries was lower than in blueberries and cranberries. Gooseberries can be used as a new source for food consumption and supplementation based on their antioxidant and binding properties. 3D fluorescence spectroscopy and FTIR spectroscopy can be applied as additional analytical tools for rapid estimation of the quality of different food products.  相似文献   

15.
The characteristics of a novel acid protease from a fusant F76 were comparatively evaluated with those from its progenitors Aspergillus oryzae HN3042 and A. niger CICC2377. The UV spectra of these three acid proteases were similar, but fluorescence spectra were different. The acid protease from F76 contained 7.1% α‐helix, 39.4% β‐sheet, 24.7% β‐turn and 32% aperiodic coil, unlike those from its progenitors. The acid protease from F76 was active in the temperature range of 35–55 °C with the optimum temperature of 40 °C and was stable in the pH range of 2.5–6.5 with the optimum pH of 3.5, while those values from A. oryzae HN3042 and A. niger CICC2377 were 45 °C, 4.0 and 40 °C, 3.5, respectively. The kinetic parameters of the acid protease from F76 were different from its progenitors and the Michaelis constant, maximum velocity, activation energy, and attenuation index were 0.96 mg mL?1, 135.14 μmol min?1 mg?1, 64.11 kJ mol?1 and 0.59, respectively.  相似文献   

16.
Stability and physicochemical properties of model sauces containing 2.5 wt% fresh beef meat (related to raw material), 30 wt% rapeseed oil and native waxy maize starch (WMS) or potato starch (PS) at concentrations ranged from 0.5 to 4.0 wt% were assessed. Sauces thickened with WMS showed a significantly (< 0.05) higher stability than respective ones made with PS. All studied systems exhibited non‐Newtonian, pseudoplastic behaviour. The Ostwald–de Waele and Herschel–Bulkley models were used to describe the flow properties of model sauces. In comparison with control sample (prepared without starch), addition of this polysaccharide (WMS or PS) increased consistency index, yield stress and apparent viscosity and decreased flow behaviour index of model sauces. The Arrhenius equation was used to determine the effects of temperature (20–50 °C) on the apparent viscosity. The activation energy values were in ranges 7.66–10.59 kJ mol?1 and 8.87–11.82 kJ mol?1 in sauces prepared with WMS and PS, respectively. The instrumentally detected changes in consistency and whiteness of model sauces were found, which may be used as the good predictors of the perceived sensory consistency and whiteness.  相似文献   

17.
The effects of temperature and pretreatment on drying kinetics and thermal degradation of phytonutrients present in pomegranate arils were investigated. The arils were divided into two groups, and half of the samples were pretreated by dipping into 80 °C hot water for 2 min. The drying process was conducted in the vacuum drier at the temperatures of 55, 65 and 75 °C. The fastest drying was completed at 75 °C after pretreatment of the samples. The highest anthocyanin–phenolic compound contents and antioxidant capacity were detected in the arils dried at 55 °C. Seven thin‐layer drying models were used to predict drying curves, and Arrhenius and Eyring–Polanyi models were employed to predict phytonutrient degradation kinetics. Activation energy for drying was 24.26 kJ mol?1 for pretreated samples and 31.54 kJ mol?1 for untreated samples. Effective moisture diffusivities were ranged from 1.43 × 10?9 to 6.03 × 10?9 m2 s?1.  相似文献   

18.
The changes of bioactive components and antioxidant activity of quince nectar were determined during 9 months of storage at 5, 20, 30 and 40 °C. The amount of total phenolics, flavonoids and antioxidant activity was significantly declined during storage at all temperatures. Loss of L‐ascorbic acid at 5, 20, 30 and 40 °C was 32.08%, 43.69%, 65.21% and 88.82%, respectively. L‐ascorbic acid degradation was in accordance with the first‐order reaction kinetics, and activation energy was found as 43.65 kJ mol?1. After 9 months of storage, Hydroxymethylfurfural (HMF) contents of quince nectars were 15.01, 16.64, 21.69 and 57.89 mg kg?1 at 5, 20, 30 and 40 °C, respectively. HMF accumulation fitted a zero‐order kinetic model, and activation energy was found as 88.30 kJ mol?1. A significant correlation was found among L‐ascorbic acid, total phenolics, flavonoids and antioxidant activity.  相似文献   

19.
The release kinetics of nisin from poly(butylene adipate-co-terephthalate) (PBAT) to distilled water was studied at of 5.6, 22 and 40 °C. The release kinetics of nisin from PBAT film was described using Fick’s second law of diffusion, partition coefficient, and Weibull model. The diffusion coefficients (D) determined were 0.93, 2.29, and 5.78 × 10−10 cm2/s at 5.6, 22, and 40 °C, respectively. The partition coefficients (K) calculated were 0.84, 3.89, and 5.2 × 103 at 5.6, 22, and 40 °C, respectively. The nisin release data at selected temperatures were fitted with the Weibull model (R2 > 0.97) with b and n values ranging from 0.02 to 0.98 and from 0.28 to 0.45, respectively. The temperature dependence of D, K, and Weibull model parameter b was modeled using the Arrhenius equation giving values of activation energy (Ea) of 38.3 kJ mol−1 (for D), 38.5 kJ mol−1 (for K), and 79.5 kJ mol−1 (for b).  相似文献   

20.
The experiments were conducted on pear slices with thickness of 5 mm at temperatures of 50, 57, 64 and 71 °C with an air velocity of 2.0 m s?1. Prior to drying, pear slices were pretreated with citric acid solution (0.5% w/w, 1 min, 20 °C) or blanched in hot water (1 min, 85 °C). Also, the untreated samples were dried as control. The shortest drying time of pear slices was obtained with pretreatment with citric acid solution. It was observed that whole drying process of pear slices took place in a falling rate period. Four mathematical models were tested to fit drying data of pear slices. According to the statistical criteria (R2, χ2 and RMSE), the Midilli et al. model was found to be the best model to describe the drying behaviour of pear slices. The effective diffusivity of moisture transfer during drying process varied between 8.56 × 10?11 and 2.25 × 10?10 m2 s?1, while the activation energy of moisture diffusion in pear slices was found to be 34.95–41.00 kJ mol?1.  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