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 625 毫秒
1.
This work describes the optimization of a cloud point extraction (CPE) method for casein proteins from cow milk samples. To promote phase separation, polyoxyethylene(8) isooctylphenyl ether (Triton® X-114) and sodium chloride (NaCl) were used as nonionic surfactant and electrolyte, respectively. Using multivariate studies, four major CPE variables were evaluated: Triton® X-114 concentration, sample volume, NaCl concentration, and pH. The results show that surfactant concentration and sample volume were the main variable affecting the CPE process, with the following optimized parameters: 1% (w/v) Triton® X-114 concentration, 50 μL of sample volume, 6% (w/v) NaCl concentration and extractions carried out at pH 7.0. At these conditions, 923 ± 66 and 67 ± 2 μg mL−1 of total protein were found in the surfactant-rich and surfactant-poor phases, respectively. Finally, matrix-assisted laser desorption ionization-time of flight mass spectrometry (MALDI-TOF MS) was then used to evaluate those target proteins (s1-casein, s2-casein and β-casein) separation as well as to check the efficiency of the extraction procedure, making a fingerprint of those target proteins possible.  相似文献   

2.
The effect of spray drying and reconstitution has been studied for oil-in-water emulsions (20.6% maltodextrin, 20% soybean oil, 2.4% protein, 0.13 M NaCl, pH 6.7) with varying ratios of sodium caseinate and whey protein, but with equal size distribution (d32=0.77 μm). When the concentration of sodium caseinate in the emulsion was high enough to entirely cover the oil–water interface, the particle size distribution was hardly affected by spray drying and reconstitution. However, for emulsions of which the total protein consisted of more than 70% whey protein, spray drying resulted in a strong increase of the droplet size distribution. The adsorbed amount of protein ranged from 3 mg m−2 for casein-stabilised emulsions to 4 mg m−2 for whey protein-stabilised emulsions with a maximum of 4.2 mg m−2 for emulsions containing 80% whey protein on total protein, which means that for all these emulsions about one quarter of the available protein was adsorbed at the oil–water interface. The adsorbed amount of protein was hardly affected by spray drying. After emulsion preparation casein proteins adsorbed preferentially at the oil–water interface. As a result of spray drying, the relative amount of β-lactoglobulin in the adsorbed layer increased strongly at the expense of s1-casein and β-casein. Percentages of s2-casein and κ-casein in the adsorbed layer remained largely unchanged. The changes in the protein composition of the adsorbed layer as a result of spray drying and reconstitution were the largest when beforehand hardly any whey protein was present in the adsorbed layer and hardly any sodium caseinate in the aqueous phase. Apparently, during spray drying conditions have been such that β-lactoglobulin could unfold, aggregate, and react with other cystein-containing proteins changing the particle size distribution of the emulsions and the composition of the adsorbed layer. It seemed, however, that non-adsorbed sodium caseinate in some way was able to protect the adsorbed casein proteins from being displaced by aggregating whey protein.  相似文献   

3.
The calcium phosphate which corresponds to the formula Ca3(PO)4 · nH2O (2<n<3) was isolated from solutions with Ca/P molar ratio 0.2 and pH 7. The compound was characterised by chemical and thermogravimetric analyses, Fourier-transform infrared (FTIR) spectroscopy and X-ray diffraction. The FTIR spectra were compared with spectra of β-tricalcium phosphate (β-TCP) in the atlases for analysis of urinary calculi and other literature data.  相似文献   

4.
A polyacrylamide gel crosslinked with allyl-β-cyclodextrin can be used repeatedly for several weeks for the separation of DNA fragments, since bubbles are not generated during a run. Allyl-β-cyclodextrin can easily be synthesized in one step from allylglycidylether and β-cyclodextrin. The plate numbers for DNA fragments, up to about 1500 bp, are high: for the separation of pBR322/HaeIII fragments they were in the range 450 000–1 600 000 m−1. The resolution was almost independent of the concentration of the crosslinker (allyl-β-cyclodextrin) — in sharp contrast to gels crosslinked with N,N′-methylenebisacrylamide.  相似文献   

5.
An intensive multi-disciplinary research effort is underway at Wayne State University to synthesize and characterize magnetic nanoparticles in a biocompatible matrix for biomedical applications. The particular system being studied consists of 3–10 nm γ-Fe2O3 nanoparticles in an alginate matrix, which is being studied for applications in targeted drug delivery, as a magnetic-resonance imaging (MRI) contrast agent, and for hyperthermic treatments of malignant tumors. In the present work we report on our efforts to determine if laser-induced breakdown spectroscopy (LIBS) can offer a more accurate and substantially faster determination of iron content in such nanoparticle-containing materials than competing technologies such as inductively-coupled plasma (ICP). Standardized samples of -Fe2O3 nanoparticles (5–25 nm diameter) and silver micropowder (2–3.5 μm diameter) were created with thirteen precisely known concentrations and pressed hydraulically to create solid “pellets” for LIBS analysis. The ratio of the intensity of an Fe(I) emission line at 371.994 nm to that of an Ag(I) line at 328.069 nm was used to create a calibration curve exhibiting an exponential dependence on Fe mass fraction. Using this curve, an “unknown” γ-Fe2O3/alginate/silver pellet was tested, leading to a measurement of the mass fraction of Fe in the nanoparticle/alginate matrix of 51 ± 3 wt.%, which is in very good agreement with expectations and previous determinations of its iron concentration.  相似文献   

6.
The dielectric permittivity tensor components, εII and ε, in the nematic phase of 6CB (4-n-hexyl-4'-cyanobiphenyl) were measured in the pressure range 0.1-130 MPa and the temperature range 12-58°C. The dielectric anisotropy, Δε(p, V, T) = εII - ε, was analysed in isothermal, isobaric and isochoric conditions taking into account the pVT data and the well known Maier and Meier equation. On that basis the nematic order parameter S(p, V, T) was determined. This was used to calculate the parameter Γ relating the interaction potential with the volume (density). Its value Γ = 4.1 agrees very well with other estimates.  相似文献   

7.
The effects of external stimuli such as pH of the buffer solution, ionic strength, temperature and the amount of poly-electrolyte monomer in the hydrogel system on the Bovine Serum Albumin (BSA) adsorption capacity of poly(acrylamide/maleic acid) [P(AAm/MA)] hydrogels were investigated. Poly-electrolyte P(AAm/MA) hydrogels with varying compositions were prepared by irradiating acrylamide/maleic acid/water mixtures with γ rays at ambient temperature. Langmuir type adsorption isotherms were observed for all prepared hydrogels. Increase of ionic strength of the buffer solution from 0.01 to 0.1 mol dm−3 decreased the adsorption capacity of hydrogels and zero adsorption was observed in the presence of 0.1 mol dm−3 Na+ and Ca2+ ion in the adsorption medium. The adsorption capacity of hydrogels was found to increase from 0 to 120 mg BSA/g dry gel, by changing external stimuli and hydrogel composition.  相似文献   

8.
Milk fat globule membrane (MFGM) lipids have been studied in the presence and absence of proteins β-lactoglobulin and β-casein. The aim of this study was to relate the self-assembly structure, e.g. vesicles, formed in aqueous dispersions of MFGM lipids to the lipid composition, electrolyte composition as well as the effect of added milk proteins, i.e. β-lactoglobulin and β-casein. For this purpose, vesicles of phospholipid mixtures, containing dioleoylphosphatidylcholine (DOPC), sphingomyelin (SM), dioleoylphosphatidylethanolamine (DOPE), phosphatidylinositol (PI) and dioleoylphosphatidylserine (DOPS) at composition corresponding to that of the MFGM, were prepared by extrusion. The morphology of the formed structures of different sample compositions was studied with cryogenic transmission electron microscopy (Cryo-TEM). Mixtures of membrane lipid with a composition (e.g. 80% DOPE, 12% DOPC and 8% SM) that at high lipid content give liquid crystalline phases at the boundary of lamellar to reversed hexagonal phase rather formed microtubular structures than vesicles at high water content. A large proportion of multilamellar vesicles is formed in buffer and divalent salts than in pure water. A small increase in the interlayer spacing of the multilamellar vesicle was observed in the presence of β-casein.  相似文献   

9.
A structural study of odd-numbered n-alkane (Cn) binary mixtures (C21 : C23) was carried out on powder samples using a Guinier-de Wolff camera with increasing concentration of n-C23 at 293 K.

Despite the reports in the literature, these molecular alloys do not form an orthorhombic continuous homogeneous solid solution to C21 from C23 at “low temperature”. Instead, as already observed in two even-numbered Cn systems, X-ray diffraction results show the existence of seven solid solutions as the molar concentration of C23 increases: four terminal solid solutions, denoted β0(C210(C23), isostructural with the “low temperature” phase of pure C21 and C23 (Pbcm), β′0(C21) and β′0(C23), identical to the phase β′0 which appears in pure C23 above the δ transition, and three orthorhombic intermediate solid solutions, designated β″1, β′1 and β″2.

On the basis of powder X-ray photographs, the phases β″1 and β″2 (C21 : C23) are indistinguishable, and they are isostructural with the intermediate solid solution β″ of the even-numbered Cn binary systems (C22 : C24) and (C24 : C26). The phase β′1(C21 : C23) is also isostructural with the two indistinguishable intermediate solid solutions β′1 and β′2 of the molecular alloys (C22 : C24) and (24 : C26).

From this study and our other laboratory results, the sequences of appearance of the solid solutions and the structural identities between these phases are established at “low temperature” for all the binary molecular alloys of consecutive Cn (odd-odd, even-even or odd-even: 19 < n < 27) when increasing the solute concentration.  相似文献   


10.
Using N3 species as specific electron acceptor a defined ascorbate radical: AH↔A+H+max=360 nm, =3400 dm3 mol−1 cm−1) is observed. The attack of DMSO+ on vit.E results in a vit.E radical (k=1×109 dm3 mol−1 s−1; λmax=425 nm, =2400 dm3 mol−1 cm−1; 2k=4.7×108 dm3 mol−1 s−1). Vit.E-acetate leads to the formation of a radical cation (vit.E-ac+). β-carotene reacts also with DMSO+ forming a radical cation, β-car+ (k=1.75×108 dm3 mol−1 s−1; λmax=942 nm, =14 600 dm3 mol−1 cm−1), which probably leads to the formation of a dimer radical cation, (β-car)+2 (k=2.5×107 dm3 mol−1 s−1).

Using E.coli bacteria (AB1157) as a model system in vitro it was found that all three vitamins are rather efficient radiation protecting agents. They can also increase the activity of cytostatica, e.g., mitomycin C (MMC), by electron transfer process. The mixture of vit.E-ac and β-car acts contradictory, but adding vit.C to it a strong cooperative enhancement of the MMC activity is observed once again. A relationship between the pulse radiolysis and the radiation biological data is found and discussed. A possible explanation of the previously reported trials concerning the role of vit.E and β-car on the increased occurence of lung and other types of cancer in smokers and drinkers is presented.  相似文献   


11.
Rapid leaching procedures by Pressurized Liquid Extraction (PLE) have been developed for As, Cd, Cr, Ni and Pb leaching from environmental matrices (marine sediment and soil samples). The Pressurized Liquid Extraction is completed after 16 min. The released elements by acetic acid Pressurized Liquid Extraction have been evaluated by inductively coupled plasma-optical emission spectrometry. The optimum multi-element leaching conditions when using 5.0 ml stainless steel extraction cells, were: acetic acid concentration 8.0 M, extraction temperature 100 °C, pressure 1500 psi, static time 5 min, flush solvent 60%, two extraction steps and 0.50 g of diatomaceous earth as dispersing agent (diatomaceous earth mass/sample mass ratio of 2). Results have showed that high acetic acid concentrations and high extraction temperatures increase the metal leaching efficiency. Limits of detection (between 0.12 and 0.5 μg g− 1) and repeatability of the over-all procedure (around 6.0%) were assessed. Finally, accuracy was studied by analyzing PACS-2 (marine sediment), GBW-07409 (soil), IRANT-12-1-07 (cambisol soil) and IRANT-12-1-08 (luvisol soil) certified reference materials (CRMs). These certified reference materials offer certified concentrations ranges between 2.9 and 26.2 μg g− 1 for As, from 0.068 to 2.85 μg g− 1 for Cd, between 26.4 and 90.7 μg g− 1 for Cr, from 9.3 to 40.0 μg g− 1 for Ni and between 16.3 and 183.0 μg g− 1 for Pb. Recoveries after analysis were between 95.7 and 105.1% for As, 96.2% for Cd, 95.2 and 100.6% for Cr, 95.7 and 103% for Ni and 94.2 and 105.5% for Pb.  相似文献   

12.
Two alkyl (1b and 1c) and four fluoroalkyl derivatives (1d-1f) of 4-arylbutyric acid (1c, 1d and 1e) and 4-arylbutanol (1b, 1f and 1g) [aryl = 2',3'-difluoro-4'-(2-(E-4-pentylcyclohexyl)ethyl)-biphenyl-1-yl] were prepared and investigated in the pure form as nematic materials (1b and 1c) and as additives to a ferroelectric liquid crystal (FLC) host (1d-1f). A comparison of 1b and 1c with the decyl analogue 1a demonstrates the effect of terminal chain modification on thermal and electro-optical properties. The substitution of the -CH2O- (1b) or -COO- (1c) for -CH2CH2- in 1a destabilized the N and SmA phases or completely eliminated the smectic behaviour (1c). Dielectric analysis revealed that the chain modification increased the negative Δε, reduced elastic constant K 11 and moderately decreased rotational viscosity γ1. The temperature dependence of the key electro-optical parameters was analysed for ester 2, the methyl analogue of 1c, which exhibits a 45 K wide N phase. All four fluoroalkyl derivatives 1d-1f showed enantiotropic SmA phases and 1e also exhibited a monotropic SmC phase. Solutions of 1d-1f in a FLC host (0.2 mmol g-1) increased the tilt angle Θ (up to 45° for 1g), reduced rotational viscosity γ1 and the risetime τ. The most dramatic changes were observed for 1g, which contains 15 fluorine atoms.  相似文献   

13.
We measured the dynamic (DLS) and static (SLS) light scattering behaviour of β-casein solutions in a 25 mM Na phosphate buffer at neutral pH as a function of temperature. At low temperatures (0 °C) β-casein is predominantly in a monomer state. With rising temperature micelles are formed with a (concentration-dependent) transition temperature in the range 15–30 °C. The transition is accompanied by a clear positive excess heat capacity. In DLS we observe two relaxation modes. The fast mode is attributed to the diffusive motion of the micelles and leads to a hydrodynamic radius of about 12 nm. The slow mode cannot be attributed to ‘physical’ particles. It is attributed to polydispersity or equivalently to long-range concentration fluctuations as proposed by Leclerc and Calmettes [15 and 16]. From SLS measurements we obtained the molecular mass and divided by the mass of a monomer (24 kDa) it gives the micellisation number, which seems to level off to about 30 at 40 °C. The measured micellisation number is predicted quite satisfactorily from a thermodynamic model for the calorimetric data as developed by Mikheeva et al. [26] and based on the shell model of Kegeles [24 and 25].  相似文献   

14.
Conditions that facilitate high Ps formation and interactions that modify the o-Ps lifetime were investigated by positron annihilation techniques in silicalite-1 and various Y-zeolites. Long lifetimes, up to 135 ns, and o-Ps fractions as high as 40% were found. The influence of heat treatment (in the range of 90–520 K), capillary condensation of N2 (0.085 MPa) and correlation with water removal was examined in Y-zeolites and in silicalite-1, respectively. The latter matrix was also studied in the presence of liquids. In various samples unexpected features were found (peculiar changes in the 2γ/3γ ratio, inversion in the trend of lifetime variations, disappearance of specific components, etc.), denoting the complexity of the processes governing the fate of e+ and Ps.  相似文献   

15.
Interfacial rheological properties and their suitability for foam production and stability of two vegetable proteins were studied and compared to β-casein. Proteins used ranged from flexible to rigid/globular in the order of β-casein, gliadin and soy glycinin. Experiments were performed at pH 6.7. Network forming properties were characterised by the surface dilational modulus (determined with the ring trough) and the critical falling film length (Lstill) at which a stagnant protein film will break. Gliadin had the highest dilational modulus, followed by glycinin and β-casein, whereas glycinin formed the strongest film against fracture in the overflowing cylinder. The rate of decrease in the surface tension was studied at the air–water (Wilhelmy plate method) and the oil–water interface (bursting membrane) and the dynamic surface tension during compression and expansion in the caterpillar. Gliadin had the lowest equilibrium interfacial tensions and β-casein the lowest dynamic surface tension during expansion. Hardly any foam could be formed at a concentration of 0.1 g/l by shaking. At a concentration of 1.4 g/l most foam was formed by β-casein, followed by gliadin and glycinin. It seems that in the first place the rate of adsorption is important for foam formation. For the vegetable proteins, adsorption was slow. This resulted in lower foamability, especially for glycinin.  相似文献   

16.
The complex (μ-H)5Os3Re(CO)12 crystallizes in the centrosymmetric hexagonal space group P63/m (C26h; No. 176) with a 19.087(5), c 10.963(1) Å, V 3459(3) Å3, and Z = 6. Diffraction data were collected on a Syntex P21 automated four-circle diffractometer (Mo-K radiation, 2θ = 4.5–45.0°) and the structure was refined to RF = 7.9% for all 1480 unique reflections (RF = 5.4% for those 1007 data with ¦Fo¦ > 6σ(¦Fo¦)). The molecule contains a tetrahedral core of metal atoms each associated with three terminal carbonyl ligands. It is bisected by a crystallographic mirror plane. Although the hydride ligands were not located, a consideration of metal-metal distances allows the distinction between osmium and rhenium atoms and suggests that the structure is subject to a subtle form of two-fold disorder.  相似文献   

17.
Milk is a complex colloidal suspension of proteins, inorganic materials and lipids. Of the proteins, caseins are present in the highest concentrations, and are themselves organised into a complex structure termed the casein micelle. The remarkable stability of the milk towards pasteurisation, sterilisation and dehydration is directly related to the stability of the casein micelle, which is in turn related to its surface components, notably κ-casein.

In this study, a surface force apparatus has been used to measure interactions between κ-casein surfaces, with a view to determining the forces involved in the stabilisation of the casein micelles. The observed interaction on a first compression is attractive, commencing at a protein surface separation of about 40 nm. This attraction is weak, around 150 μN m−1, and may be hydrophobic in origin. The thickness of the κ-casein layer is 9.8 nm. On separation of the surfaces, no attraction is noted, only a short-range steric interaction being seen, indicating that some configurational change of the protein is occurring.  相似文献   


18.
The adducts of O2 and SO2 with trans-MeOIr(CO)(PPh3)2 are formed in equilibria and have been characterized. Reaction of the SO2 adduct, Ir(OMe)(SO2)(CO)(PPh3)2 with dioxygen leads to the sulfato complex, Ir(Ome)(CO)(PPh3)2(SO4), the structure of which has been determined. Ir(Ome)(CO)(PPh3)2(SO4) crystallizes in the monoclinic system with a 11.958(2), b 14.163(3), c 12.231(2) Å, β 118.365(12)°, V 1822.7(6) Å3 and Z = 2. Diffraction data for 2θ = 4.5–45.0° (Mo-K) were collected with a Syntex P21 diffractometer and the structure was solved (assuming space group P21/m and an unpleasant 2-fold disordered model) and refined to R = 4.8% for all 2512 independent data (R = 3.5% for those 2042 data with ¦FO¦ > 6σ(¦F¦)). The iridium(III) atom has a distorted octahedral coordination sphere with trans PPh3 ligands and a cis-chelating bidentate O,O′-SO4 group; the structure is completed by mutually cis OMe and CO ligands.  相似文献   

19.
Li X  Xiong Z  Ying X  Cui L  Zhu W  Li F 《Analytica chimica acta》2006,580(2):170-180
A rapid ultra-performance liquid chromatography-electrospray ionization tandem mass spectrometric (UPLC–ESI-MS/MS) method was developed for the qualitative and quantitative determination of the constituents of the flower of Trollius ledibouri Reichb. The analysis was performed on an AcQuity UPLC™ BEH C18 column using gradient elution with a mobile phase of 0.1% acetic acid and acetonitrile over 20 min. A tandem quadrupole spectrometer operating in either full scan mode or in MS/MS mode for multiple reaction monitoring (MRM) was used for the qualitative and quantitative analysis of the constituents, respectively. According to the mass spectrometric fragmentation mechanism and UPLC–ESI-MS/MS data, the chemical structures of 15 constituents of the flower of T. ledibouri Reichb. were identified on-line without time-consuming isolation and four of them, 2″-O-β-l-galactopyranosylorientin, 2″-O-β-arabinopyranosylorientin, orientin and vitexin, were quantified. The limits of quantification of these four flavonoids were 540, 321, 515 and 220 μg g−1 plant material, respectively. Four commercial samples from different sources were analyzed. The UPLC–ESI-MS/MS method for analyzing the constituents can be used to evaluate the quality of the flower of T. ledibouri Reichb.  相似文献   

20.
The concentration dependences of dielectric properties measured at 105 Hz and 106 Hz are reported for aqueous solutions of hydroxypropyl cellulose. Phase behaviour of the solutions was also observed with a polarizing optical microscope. For solutions with concentrations well above 40 wt %, polydomain textures, including the banded texture, were observed after a prehistory of deformation. No significant discontinuous changes in the dielectric constant, εr', and loss factor, εr', were found at the concentrations around the onset of the isotropic-cholesteric phase transition and in the biphasic region. In contrast, the steeper changes in εr' and εr' were found at the critical concentration for the fully developed cholesteric phase transition with the polydomain textures.  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