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.
In this study, gelatin blended with arabic gum microcapsules containing camphor oil with added polystyrene were fabricated by a compound coacervation method. The parameters of oil/wall volume ratio, emulsification stirring speed, concentration of cross-linking agent, treated time and oil release properties were investigated. In order to improve the constant release effect of camphor oil, oil-soluble polystyrene (PS) was used as a sustained release agent. The camphor oil release curves were expressed by the exponential equation: ψ(t) = Ceq(1–et/τ), where ψ(t) represent the variant of camphor oil concentration in the operation environment, Ceq as the equilibrium state, t as the release time and τ as time constant. Ceq and τ are significant factors pertaining to the camphor oil release properties. The results indicated that, for the microcapsules, the optimal oil/wall volume ratio was 0.75 to achieve the encapsulation efficiency of 99.6 wt.%. The average particle size were 294.7 ± 14.2 μm, 167.2 ± 11.2 μm, 85.7 ± 8.7 μm at the homogenization stirring speed of 500, 1000, and 2000 rpm, respectively. The effect of sustained oil release will increase whereas the stirring speed decreases and the concentration of glutaraldehyde (GA) and treated time increases. Along with the increasing of quantity of polystyrene added, Ceq decreased and τ increased, indicating that the sustained oil release amount and the release rate depend on the quantity of PS considerably.  相似文献   

2.
A metal-organic complex, which has the potential property of absorbing gases, [LaCu6(μ-OH)3(Gly)6im6](ClO4)6 was synthesized through the self-assembly of La3+, Cu2+, glycine (Gly) and imidazole (Im) in aqueous solution and characterized by IR, element analysis and powder XRD. The molar heat capacity, Cp,m, was measured from T = 80 to 390 K with an automated adiabatic calorimeter. The thermodynamic functions [HT − H298.15] and [ST − S298.15] were derived from the heat capacity data with temperature interval of 5 K. The thermal stability of the complex was investigated by differential scanning calorimetry (DSC).  相似文献   

3.
A new 1.75 μm infrared emission transition of Y2O3:Er3+ is assigned to the 4S3/2 → 4I9/2 transition of Er3+ ions situated at the C2 sites of cubic RE2O3 (RE = Y, Gd, Lu). The intensities of features in the 1.54 μm 4I15/24I13/2 absorption transition due to Er3+ at S6 and C2 sites are consistent with the site occupation ratio and the relative magnetic dipole–electric dipole intensity contributions of Er3+ at the different sites. The 1.54 μm emission lines are predominantly from Er3+ ions at C2 sites. The different behaviours of the emission intensities 1.75 and 1.54 μm groups with change in Er3+ dopant ion concentration, preparation technique, Yb3+ co-doping, temperature change and different excitation line are rationalized.  相似文献   

4.
Surface properties of foam films formed from aqueous dispersions of dipalmitoyl-phosphatidylcholine (DPPC) and from solutions of a phospholipid fraction of lung surfactant (TPL) are studied employing the foam film method. Experiments are carried out within a wide range of NaCl concentrations (Cel) and the ranges of Cel determining formation of common films (CF), common black films (CBF) and bilayer Newton black films (NBF) are found. The thickness (h) of the CF and CBF decreases with the increase of Cel until the critical electrolyte concentrations (Cel, cr) is reached. The determined Cel, cr that characterize the transition to NBF show that Cel, cr of the TPL films is an order of magnitude higher than that of the DPPC films. The measured h of the TPL films is higher than that of the DPPC films in the whole Cel range. Besides, only the h(Cel) curve of the DPPC films outlines a metastable Cel range where both CF and NBF are obtained. Both the h(Cel) curves and the direct measurements of the disjoining pressure isotherms of the DPPC films (Π(h) isotherms) demonstrate the role of electrostatic repulsive forces for the stability of the phospholipid films The obtained results are compared with the DLVO theory equations and the evaluated potentials of the diffuse electric layer φ0  20 mV for the DPPC films and φ0  100 mV for TPL films show the strong effect of the charged phospholipids in the TPL mixture on the electric properties at the film interfaces.  相似文献   

5.
CoAl-MCM-41 (X) catalysts with X = nSi/(nCo + nAl) various ratios were synthesized and ethylation of phenol with ethanol was studied in vapor-phase at temperatures between 250 and 450 °C. The products obtained were O-alkylated product (ethyl phenyl ether), C-alkylated products (2-ethylphenol and 4-ethylphenol), and C-/O-alkylated products (ethyl ethylphenyl ether). The phenol conversion increased significantly with reaction temperature over all the catalysts. The activity of the catalysts followed the order CoAl-MCM-41 (20) > CoAl-MCM-41 (50) > CoAl-MCM-41 (80). Selectivity between the C-alkylation and the O-alkylation depended on the factors such as acidity of the catalyst and the reaction temperature. CoAl-MCM-41 (20) catalyst displayed a phenol conversion of 40% and a selectivity of more than 80% for 2-ethylphenol under the optimized reaction condition. The ethanol to phenol ratios and the reactant flow rate are also influential for both activity and selectivity of CoAl-MCM-41 catalysts.  相似文献   

6.
Effects of the surfactant concentration Cd and the NaCl concentration Cs on the electrophoretic mobilities U of the well-characterized polymer-like micelles have been investigated by the electrophoretic light scattering, using tetradecyldimethylamine oxide hemihydrochloride (C14DMAO·1/2HCl). At the high ionic strength of 0.1 mol kg−1 NaCl, the electrophoretic mobilities were independent of Cd (5 mM < Cd < 100 mM), despite the concentration-dependent micelle growth of the polymer-like micelles. This suggests that the electrophoretic mobility of the polymer-like micelle at high ionic strengths is independent of the contour length (i.e., the molecular weight), as found on linear polyelectrolytes. Somewhat surprisingly, the entanglements of the polymer-like micelles gave small effect on the electrophoretic mobilities in the examined range of the surfactant concentration above an overlap concentration. The mobilities of the polymer-like micelle decreased with √Cs in a single exponential manner in the range of Cs from 0.02 to 0.3 mol kg−1. It is suggested that the cylinder model can be applied to the electrophoretic mobilities of the polymer-like micelles at high ionic strengths (i.e. a free-draining behavior), since the persistence length of the polymer-like micelle (20 nm) is much larger than the Debye length at high ionic strength.  相似文献   

7.
The structures of 3,3′-dicarbometoxy-2,2′-bipyridine (dcmbpy) complexes with copper(II) and silver(I) cations have been determined using single crystal X-ray-diffraction. The crystals of Cu(dcmbpy)Cl2 are monoclinic, C2/c, a = 16.966(3), b = 18.373(3), c = 13.154(2) Å, β = 126.543(3)°. The crystals of Ag(dcmbpy)NO3 · H2O are also monoclinic, C2/c, a = 16.7547(13), b = 11.0922(9), c = 18.7789(18) Å, β = 100.228(7)°. The results have been compared with the literature data on the complexes of dcmbpy and its precursors: 2,2′-bipyridine (bpy) and 3,3′-dicarboxy-2,2′-bipyridine (dcbpy). Two types of complexes of 3,3′-carboxy derivatives of bpy are distinguished: (1) with metal atom bonded to two N atoms of the same molecule and (2) with metal atom bonded to two N atoms of two different molecules. The Cu(dcmbpy)Cl2 complex belongs to the first type, whereas Ag(dcmbpy)NO3 · H2O belongs to the second type.  相似文献   

8.
Ferdousi BN  Islam MM  Okajima T  Ohsaka T 《Talanta》2008,74(5):1355-1362
We successfully determined the molecular structure of peroxycitric acid (PCA) coexisting in the aqueous equilibrium mixture with citric acid (CA; 1,2,3-tricarboxylic-2-hydroxy propane) and hydrogen peroxide (H2O2) by a combined use of reversed-phase HPLC (RP-HPLC), potentiometric, hydrodynamic chronocoulometric (HCC) and electrospray ionization mass spectroscopic (ESI-MS) methods. Firstly, the RP-HPLC was employed to separate CA, PCA and H2O2 coexisting in the equilibrium mixture and the concentration of CA consumed (ΔCCA) in the formation of PCA that was evidenced to be fairly stable during the RP-HPLC measurement was quantitatively measured based on the standard calibration curve of CA. Secondly, the total oxidant concentration (COx) corresponding to peroxycarboxylic (–COOOH) group in PCA in the equilibrium mixture was determined using potentiometric measurement. The ratio of COxCCA was found to be 1.07, which indicates that only one –COOH group in CA molecule is oxidized to the corresponding –COOOH group in PCA molecule. Thirdly, using the HCC technique the diffusion coefficient of PCA, which could be electroreduced at a more positive potential by 1.0 V than the coexisting H2O2, was independently measured as 0.3 × 10−5 cm2 s−1 and at the same time, by considering ΔCCA as the concentration of PCA, the number of electrons (n) required for the reduction of PCA was determined to be 2. The result obtained from RP-HPLC and HCC, i.e., n = 2 which is equivalent to one –COOOH group in PCA, is in agreement with that obtained from the combination of RP-HPLC and potentiometric measurements. Finally, the structure of PCA was proposed to contain one –COOOH group with a molecular mass of 208 confirmed by negative ion ESI-MS method. A probable molecular structure of PCA was discussed.  相似文献   

9.
We studied the effects of the degree of ionization() and the surfactant concentration (Cd) on the micelle–vesicle transition in salt-free oleyldimethylamine oxide (OlDMAO) aqueous solutions by the dynamic light scattering (DLS), the hydrogen ion titration, the small angle neutron scattering (SANS), the electrophoretic light scattering (ELS) and viscoelastic measurements. From the study of ionization effects, the micelle–vesicle transition was recognized as a change of aggregate size by the DLS measurement; however, the micelle–vesicle transition was not detected both in the ELS measurement and the hydrogen ion titration, suggesting that the electric properties of the worm-like micelles and the vesicles are very similar despite a large difference of shapes between them. From the results of the SANS, the DLS and the viscosity measurements, it was suggested that a concentration-dependent micelle–vesicle transition took place around Cp = 10 mmol kg−1 for the solutions at = 0.5. In the concentration-range 10 mmol kg−1 < Cd < 150 mmol kg−1, the micelles and the vesicles coexisted. In the concentration region (Cd = 10–50 mmol kg−1), the vesicle size increased with the surfactant concentration.  相似文献   

10.
S0 → S1 and S0 → S2 electronic transitions have been observed in UV–Visible absorption spectroscopy of 3-pyrazolyl-2-pyrazoline (PZ) in different homogeneous solvents. Radiative emissions and relaxation processes from S1 and S2 states of PZ have been resolved in water, ethylene glycol and glycerol whereas in polar aprotic and protic solvents the radiative transitions have been observed from S1 state. The S2–S1 electronic energy spacing has been calculated from the absorption maxima of the S0 → S2 transitions and fluorescence maxima of the S1 → S0 transitions. Solute–solvent interactions have been established to rationalize the photophysical modification of PZ in H-bonding solvents.  相似文献   

11.
The catalytic activity of MV2O6 and M2V2O7 type oxides prepared by the molten method (MM) for anaerobic oxidation of isobutane was studied in order to construct a system for the selective oxidation of isobutene using a thin layer reactor. Isobutene, CO and CO2 were formed by every catalyst tested. The activities for isobutene formation were CuV2O6 > ZnV2O6, NiV2O6, CoV2O6 > MgV2O6 > MnV2O6  CaV2O6. Isobutene was a major product over M2V2O7 (MM). Co2V2O7 showed the highest activity and high isobutene selectivity exceeded 90%, demonstrating that Co2V2O7 is a suitable oxide for a thin layer reactor for anaerobic oxidation of isobutane. Partial substitution of Mg by Cu in Mg2V2O7 (MM) improved the activity. It is shown by the oxidation at low O2 concentration as 2–3% that two types of oxidations occurred simultaneously: isobutene formation by the lattice oxygen ions diffused from the bulk, and CO and CO2 formation by the oxygen species derived from molecular oxygen in the gas phase.  相似文献   

12.
A chemiluminescent flow system for bromate detection, based on the reaction of bromate with sulphite in acid medium and using the steroid hydrocortisone as sensitiser, was studied. A factorial analysis strategy for the study of the effect on the system response of the experimental factors, flow rates of two pumps (Q1 — acid sulphite plus hydrocortisone aqueous solution; Q2 — carrier, water), sample injection volume (VL), reactor volume (VR), sulphite concentration (CS), hydrocortisone concentration (CH) and acid concentration (CA), was used. Screening analysis of the system performance was made using Plackett Burman designs. The system optimisation procedure was achieved by three levels three factors full factorial designs. VL and CH are the most significant factors — a quadratic CH term was also observed to be significant. The optimised system responded linearly (logarithm of the detector signal as function of the logarithm of the bromate concentration) in the concentration range between 3.6×10−7 and 5.0×10−4 M with a limit of detection of about 8.0×10−8 M (about 10 microg/l). An analysis of some interfering ions was made and it was suggested that bromide and chloride begin to quench chemiluminescence when they are in a 10-fold excess relatively to bromate concentration.  相似文献   

13.
The disruption of lipidic metabolism was considered a good candidate to explain FB1 toxicity mechanism. In the present work we investigated molecular organizational changes induced by FB1–biomembrane interaction possibly involved in mycotoxic effects.

FB1 was self-aggregated with a critical micellar concentration of 1.97 mM. FB1 (0–81.4 μM), decreased in a dose-dependent manner, the fluorescence anisotropy of TMA-DPH (from 0.349 ± 0.003 to 0.1720 ± 0.0035) in dpPC bilayers, whilst no differences were registered with DPH. At 5.6 μM in the subphase, FB1 increased the lateral surface pressure (π) of a Langmuir film to an extent that depended on the monolayer composition (ΔπdpPC:DOTAP 3:1 > ΔπdpPC:dpPA3:1 > ΔπdpPC), the molecular packing (Δπ decreased linearly as a function of the initial π) and the subphase pH (ΔπpH 2.6 > ΔπpH 7.4 and maximal π allowing the drug penetration πcut-off was 34.3 and 27.7 mN/m at pH 2.63 and 7.4, respectively). FB1 increased the surface potential of dpPC and dpPC:DOTAP monolayers and decreased that of dpPC:dpPA. This suggested that FB1 acquired different orientations and/or foldings depending on the surface electrostatics and the toxin charge state. Moreover, FB1–lipid interactions were transduced into long-range effects at the mesoscopic level affecting the lipidic self-separated lateral domains shape and density.  相似文献   


14.
Spatial structure of six β-substituted enones, with common structure R1O–CR2CH–COCF3, were R1 = C2H5, R2 = H (ETBO); R1 = R2 = CH3 (TMPO); R1 = C2H5, R2 = C6H5 (ETPO); R1 = C2H5, R2 = 4- O2NC6H4 (ETNO); R1 = C2H5, R2 = C(CH3)3 (ETDO) were investigated by 1H and 19F NMR, infrared spectroscopy and AM1 calculations. NMR spectra revealed that enones (MBO), (ETBO) and (TMPO) are exclusively (3E) isomers, whereas in (ETPO), (ETNO) and especially in (ETDO) the percentage of (3Z) isomers is significant and depends on the nature of solvents. Conformational behaviour of studied enones are determined by the rotation around of CC double bond, C–C and C–O single bonds (correspondingly trifluoroacetyl and alkoxy groups), and (EZZ) conformer being the most stable in all cases. IR spectra revealed that with the exception of (ETDO) (EZZ) conformer is most populated in all cases. Bulky substituents like phenyl or tert-butyl group at β-position of enone result in the equilibrium mainly between (EZZ) and (ZZZ) forms, whereas β-hydrogen and β-methyl substituents determine the equilibrium between (EZZ) and (EEZ) or (EZE) conformers.  相似文献   

15.
In a previous paper [Ding et al., J. Membr. Sci. 276 (2006) 232], we have investigated the performance in microfiltration of mineral suspensions of a novel filtration pilot consisting in overlapping ceramic membranes disks rotating at same speed on two parallel shafts. In this paper, we investigate a modification of this concept in which the ceramic disks of one shaft were replaced by non-permeating metal disks of same size rotating at a speed different from that of membranes. We also operated the pilot without disks on the 2nd shaft in order to eliminate membrane overlapping. When using metal disks with radial vanes, permeate fluxes were found to be 50–60% higher than those obtained in the same conditions with the previous design using only ceramic disks. By comparing permeate fluxes in different configurations, membranes on both shafts, membranes on the 1st shaft with and without metal disks on the 2nd shaft, we showed that, at a feed concentration of 200 g L−1, the effect on permeate flux J, of shear rate increment due to membrane overlapping, could be completely offset by the high concentration increase between two adjacent and overlapping membranes. Raising the ceramic disks rotation speed Nc had a larger effect on J than increasing the metal disks speed Nm. For Nc = 32.16 Hz (1930 rpm) and Nm = 2.4 Hz (144 rpm), J reached 1790 L h−1 m−2 at 310 kPa, versus 1100 L h−1 m−2 for Nc = 12.3 Hz (738 rpm) and Nm = 22.26 Hz (1336 rpm) (for the same total sum Nc + Nm). Measurements of electrical power consumed by friction on rotating disks showed that the energy spent per m3 of permeate was lowest when using metal disk with vanes rotating at low speed and ceramic disks rotating at high speed.  相似文献   

16.
The electrochemical properties of mer-[RuCl3(dppb)(4-pic)] (dppb = Ph2P(CH2)4PPh2, 4-pic = CH3C5H4N), Rupic, in CHCl3 are governed by the formation of species such as [Ru2Cl5(dppb)2], [Ru2(dppb)2Cl4(4-pic)] and trans-[RuCl2(dppb)(4-pic)2] upon the reduction of “[RuCl2(dppb)]”. The overall behavior depends on whether Rupic is immobilized in cast or Langmuir–Blodgett (LB) films, or incorporated into a carbon paste electrode (CPE). In cyclic voltammograms, one redox process appears for LB/Rupic films and CPE/Rupic, at Epa = 0.35 V, Epc = 0.25 V vs SCE, and Epa = 0.32 V, Epc = 0.24 V vs Ag/AgCl, respectively. This redox process was ascribed to the RuIII/RuII charge transfer. For cast films the redox pair was poorly defined, with Epa = 0.27 V and Epc = 0.20 V. The reason for the difference lies in the phase separation and formation of aggregates onto ITO for the cast film, in contrast to the LB film. With aggregation, the formation of species occurring in solution is impaired for Rupic in cast films. The electrochemical properties for Rupic in LB films and incorporated into CPE allowed the electrocatalytic activity of Rupic to be exploited in sensors for dopamine and ascorbic acid.  相似文献   

17.
The structure and texture characteristics of the hybrid organic–inorganic adsorbents, which were obtained by using of two-component systems of “structure-forming agent/trifunctional silane”, are compared as follows: the first component is Si(OC2H5)4 or (C2H5O)3Si–A–Si(OC2H5)3, where A = –(CH2)2– or –C6H4–; the second one is alkoxysilane with amine (–NH2, NH, –NH(CH2)2NH2) and thiol (–SH) groups. The adsorbents, derived from TEOS, have more accessible functional groups (2.6–4.2 mmol/g) than xerogels, which are based on bis(triethoxysilanes) (1.0–2.6 mmol/g). On another hand xerogels derived from bis(triethoxysilanes) have a more extended porous structure (Ssp =516–968 m2/g, Vs = 0.418–1.490 cm3/g, d = 2.5–15.0 nm) than those that are based on TEOS (Ssp = 4–631 m2/g, Vs = 0.005–1.382 cm3/g, d = 2.3–17.7 nm). The geometric dimensions of functional groups have a more essential effect on the parameters of porous structure in the case of TEOS-derived xerogels. Using solid-state NMR spectroscopy, it has been shown that in synthesis of xerogels with the use of TEOS, the molecular frame of globules is formed by structural units Qn (n = 2,3,4), and the functional groups exist as structural units of Tn (n = 2,3). The xerogels obtained with using bis(triethoxysilanes) consist only of structural units of Tn-type (n = 1,2,3).  相似文献   

18.
Excited states population distributions created by two-step 6S1/2 → 6P3/2 → 6D5/2 laser excitation in room temperature cesium vapor were quantitatively analyzed applying absorption and saturation spectroscopy. A simple method for the determination of the excited state population in a single excitation step that is based on the measurements of the saturated and unsaturated absorption coefficients was proposed and tested. It was shown that only ≈ 2% of the ground state population could be transferred to the first excited state by pumping the Doppler broadened line with a single-mode narrow-line laser. With complete saturation of the second excitation step, the population amounting to only ≈ 1% of the ground state can be eventually created in the 6D5/2 state. The fluorescence intensity emerging at 7P3/2 → 6S1/2 transition, subsequent to the radiative decay of 6D5/2 population to the 7P3/2 state, was used to assess the efficiency of the population transfer in the chosen two-step excitation scheme. The limitations imposed on the sensitivity of such resonance fluorescence detector caused by velocity-selective excitation in the first excitation step were pointed out and the way to overcome this obstacle is proposed.  相似文献   

19.
The electronic and geometrical structures of the low-energy states of 1,4,5,8-naphthalenetetracarboxylic dianhydride parent diimide (1) are studied in terms of the complete active space self-consistent field (CASSCF) method employed at different level with respect to the size and the quality of the active space. In the framework of the vibronic model based on the Franck–Condon (FC) effect the absorption and magnetic circular dichroism (MCD) spectra are studied in the excitation region corresponding to two low-energy 11Ag → 11B2u and 11Ag → 11B3u electronic transitions in diimides. In that (visible) excitation region the CASSCF computations with the 5π[4n]5π active space (i.e., the naphthalene-like π orbitals enriched by the four lone pair orbitals of the oxygen atoms) were found to reproduce very well the empirical absorption and the MCD spectra measured for the dicyclohexyl-N,N-substituted diimide (2). At the same CASSCF/5π[4n]5π level, the electronic absorption of diimides in the near UV excitation region were attributed to the 11Ag → 21B1u, 11Ag → 21B3u and 11Ag → 21B2u electronic transitions; the latter two are mostly localized on the “diimide chromophore”. For these transitions the calculated magneto-optical characteristics, such as sign pattern and intensity distribution in the MCD spectrum, were found to be consistent with that experimentally observed for the diimide 2 compound.  相似文献   

20.
Adsorption, desorption and degradation by DNase I of DNA on montmorillonite (M) and different hydroxyaluminum-M complexes (Al(OH)x-M) containing 2.5, 10.0 and 20.0 mmol coated Al/g clay (AM2.5, AM10 and AM20) were studied. The adsorption isotherms of DNA on montmorillonite and Al(OH)x-M complexes conformed to the Langmuir equation. The amount of DNA adsorbed followed the sequence of montmorillonite > AM20 > AM10 > AM2.5. A marked decrease in the adsorption of DNA on montmorillonite and Al(OH)x-M complexes was observed with the increase of pH from 4.0 to 9.0. Calcium ion significantly promoted DNA adsorption. The adsorption enthalpy of DNA on montmorillonite was endothermic, whereas that on Al(OH)x-M complexes was exothermic. The percent desorption of DNA from clays was in the order of montmorillonite > AM2.5 > AM10 > AM20, suggesting that OH–Al loading on montmorillonite surface increased the binding affinity of DNA. Fourier transform infrared (FTIR) spectra showed that the binding of DNA on AM10 and AM20 changed its conformation from the B-form to the Z-form. The presence of montmorillonite and Al(OH)x-M complexes provided protection for DNA against degradation by DNase I. The higher level of protection was found with Al(OH)x-M complexes compared to montmorillonite. The higher stability of DNA in the system of Al(OH)x-M complexes seemed to be attributed mainly to the conformational change of bound DNA and their greater adsorption capacity for DNase I. The information obtained in this study is of fundamental significance for understanding the behavior of extracellular DNA in soil environments.  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