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.
Liquid–liquid equilibrium data for {1-methyl-2-pyrrolidinone (NMP) + heterocyclic nitrogen compounds + hexadecane} systems were analytically determined at 298.15 K and atmospheric pressure using stirred and thermo-regulated cell. The experimental data were modeled with the NRTL and UNIQUAC equations. Besides, the Bachman–Brown correlation was used to ascertain the reliability of the experimental data. Additionally, excess molar volumes (VE) and deviations in the molar refractivity (ΔR) data at 298.15 K were determined for the {NMP + heterocyclic nitrogen compounds} binary systems using a digital vibrating-tube densimeter and a precision digital refractometer. The VE and ΔR data were modeled with the Redlich–Kister equation.  相似文献   

2.
Experimental values for mutual diffusion coefficients (interdiffusion coefficients) for CaCl2 aqueous solutions at 298.15 K and 310.15 K, not available in the literature, were experimentally determined between 0.005 mol dm?3 and 0.1 mol dm?3. The measurements were performed by using a conductimetric open-ended capillary cell. Values at infinitesimal concentration, D0, were also obtained. These data were analysed using the Onsager–Fuoss and Gordon models. In addition, molar conductance data were measured at 310.15 K for CaCl2 aqueous solutions at the same concentration range and the value at infinitesimal concentration determined. Afterwards, it was split into the ionic limiting values. By replacing them into the Nernst equation, diffusion coefficients at infinitesimal concentration were derived (1.335 × 10?9 m2 s?1 and 1.659 × 10?9 m2 s?1) at both temperatures (298.15 K and 310.15 K, respectively). They agree well with the extrapolated experimental ones (1.312 × 10?9 m2 s?1 and 1.613 × 10?9 m2 s?1).  相似文献   

3.
We report experimental measurements of the phase behavior of (CO2 + H2O + NaCl) and (CO2 + H2O + KCl) at temperatures from 323.15 K to 423.15 K, pressure up to 18.0 MPa, and molalities of 2.5 and 4.0 mol kg−1. The present study was made using an analytical apparatus and is the first in which coexisting vapor- and liquid-phase composition data are provided. The new measurements are compared with the available literature data for the solubility of CO2 in brines, many of which were measured with the synthetic method. Some literature data show large deviations from our results.The asymmetric (γφ) approach is used to model the phase behavior of the two systems, with the Peng–Robinson equation of state to describe the vapor phase, and the electrolyte NRTL solution model to describe the liquid phase. The model describes the mixtures in a way that preserves from our previous work on (CO2 + H2O) the values of the Henry's law constant and the partial molar volume of CO2 at infinite dilution Hou et al. [22]. The activity coefficients of CO2 in the aqueous phase are provided. Additionally, the correlation of Duan et al. [14] for the solubility of CO2 in brines is tested against our liquid-phase data.  相似文献   

4.
K and Mg substituted perovskite catalysts La1  xKxCo1  yMgyO3 (x = 0–0.4, y = 0–0.2) for soot combustion were prepared by citric acid complexation and characterized by XRD, FT-IR, SEM, TEM, EDS, H2-TPR, XPS and TG. Soot combustion was remarkably accelerated when K was introduced into LaCoO3. Then Mg was doped into the K substituted LaCoO3, soot combustion was further improved for the restrained growth of Co3O4 phase. K/Mg substitutions were responsible for enhancing activity of catalysts by improving reducibility as suggested by H2-TPR studies. Among all the catalysts, La0.6K0.4Co0.9Mg0.1O3 exhibited the highest activity.  相似文献   

5.
The fracture behavior of the notched unidirectionally solidified eutectic Al2O3/YAG composite was investigated by tensile test at room temperature to 2023 K, and the results were analyzed by the finite element method. The stress–strain behavior, notched strength and the fracture toughness were strongly dependent on temperature and displacement speed. The specimens fractured in a brittle manner at low temperatures and at high displacement speeds but in a ductile manner accompanying plastic deformation at high temperatures and at low displacement speeds. The notched strength for a given displacement speed of 10−7 m/s increased with increasing temperature from 132 MPa at 1873 K to 153 MPa at 1923 K, and then decreased to 133 and 110 MPa at 1973 and 2023 K, respectively. Also the notched strength for a given temperature of 2023 K increased with decreasing displacement speed from 67 MPa at the displacement speed of 10−5 m/s to 124 MPa at 10−6 m/s, and then decreased to 110 and 72 MPa at 10−7 and 10−8 m/s, respectively. The stress distribution and the plastic zone size ahead of the notch were calculated by a finite element method using the dependence of flow stress on temperature and displacement speed. Based on the calculation result, the experimentally observed increase in the notched strength with increasing temperature and decreasing displacement speed up to the maximum value could be accounted for by the increase in plastic zone size ahead of the notch. Also, the observed decrease in notched strength with further increasing temperature and decreasing displacement speed could be accounted for by the decrease in the stress carrying capacity of the yielded ligament.  相似文献   

6.
A visual and volume-variable high-pressure phase equilibrium analyzer was used for measuring the vapor–liquid phase boundaries of the ternary systems containing carbon dioxide and mixed solvents of dimethyl sulfoxide (DMSO) + ethanol or chloroform + methanol at temperatures from 298.15 K to 348.15 K over wide composition ranges including near critical points. Four pseudo-binary systems of carbon dioxide plus mixed solvents with constant molar ratios of DMSO/ethanol = 3/7, 5/5, 7/3, and chloroform/methanol = 1/2, were studied in this work. The critical conditions at each investigated temperature were estimated from the experimental isothermal phase boundaries by interpolation. The Peng–Robinson equation of state with the two-parameter van der Waals mixing rules was applied to calculate the phase boundaries. The experimental values were compared with the predicted results from the Peng–Robinson equation using the binary interaction parameters determined from the vapor–liquid equilibrium data of the constituent binaries. The agreement is reasonably well for carbon dioxide + chloroform + methanol, but obvious overestimations are found near the critical regions of carbon dioxide + DMSO + ethanol, especially at higher temperatures.  相似文献   

7.
The aim of this research was to investigate the phase equilibrium behavior of a system containing guaçatonga extract + ethanol + CO2 in order to help define the adequate conditions of temperature and pressure for the co-precipitation process, performed by means of supercritical anti-solvent (SAS) technique. Guaçatonga (Casearia sylvestris) is a native medicinal plant from Brazil, rich in valuable components such as β-caryophyllene, α-humulene and bicyclogermacrene. Phase equilibrium data were obtained by the static method using guaçatonga extract dissolved in ethanol (1:100, wt/wt), at temperatures ranging from 35 to 75 °C and CO2 mass content from 60 to 90 wt%. It was noticed that the system exhibited solid–vapor–liquid, solid–liquid–liquid and solid–vapor–liquid–liquid transition types and a lower critical solution temperature behavior. Phase behavior study was considered for the definition of the SAS conditions applied for the encapsulation of guaçatonga extract in the biopolymer Pluronic F127. The conditions tested ranged from 80 to 140 bar at 45 °C. At 80 bar only segregated particles of extract and the biopolymer were detected, while at 110 and 140 bar an extract encapsulation was achieved.  相似文献   

8.
Biodegradable polymers have received increased attention due to their potential applications in the medicine and food industries; in particular, poly(l-lactic acid) (PLA) is of primary importance because of its biocompatibility and resorbable features. Recently, the synthesis of this biopolymer through the enzyme-catalyzed ring-opening polymerization of l-lactic acid in a compressed fluid has been considered promising. The aim of this work was to report the phase equilibrium data (cloud points) of the l-lactic acid + (propane + ethanol) and the l-lactic acid + (carbon dioxide + ethanol) systems. The phase equilibrium experiments were conducted in a variable-volume view cell employing the static synthetic method. These experiments were conducted in the temperature range of 323.15–353.15 K and at pressures up to 25 MPa; the mass ratio of ethanol to either CO2 or propane was maintained at 1:9. The l-lactic acid + (propane + ethanol) system exhibited vapor–liquid, liquid–liquid and vapor–liquid–liquid transitions, whereas the l-lactic acid + (carbon dioxide + ethanol) system only exhibited liquid–liquid type transitions.  相似文献   

9.
The experimental solubility of dibenzofuran in near-critical and supercritical carbon dioxide and the solid–liquid–vapor (SLV) equilibrium line for the CO2 + dibenzofuran system are reported. The built in-house static view cell apparatus used in these measurements is described. The solubility of naphthalene in supercritical CO2 and the CO2 + naphthalene SLV line are also determined in order to assess the reliability and accuracy of the measurement technique. The solubility of dibenzofuran in carbon dioxide is determined at 301.3, 309.0, 319.2, 328.7 and 338.2 K in the 6–30 MPa pressure range. Solubility data are correlated using the Chrastil model and the Peng–Robinson equation of state. This equation is also used to predict the CO2 + dibenzofuran SLV line. Results show the feasibility of using supercritical CO2 to extract dibenzofuran.  相似文献   

10.
New spinel-types of S2O82 /ZnFexAl2  xO4 solid acid catalysts were prepared by sol–gel method. Their catalytic performances for the synthesis of n-butyl acetate were investigated. The catalysts were characterized by means of XRD, IR, XPS, FT-IR of adsorbed pyridine and NH3-TPD. The experimental results showed that S2O82 /ZnFexAl2  xO4 solid acid catalysts maintained the spinel structure as well as the support of ZnFexAl2  xO4. Fe3 + ions were well incorporated and highly dispersed into the spinel lattice. S2O82 /ZnFe0.15Al1.85O4 exhibited the maximum conversion of acetic acid with 98.2%. Moreover, S2O82 /ZnFe0.15Al1.85O4 showed better reusability, which remained above 72.7% conversion of acetic acid even after being used five times.  相似文献   

11.
We report here on the activity and stability of low-content praseodymium–, samarium– and gadolinium–cerium oxide catalysts for the steam reforming of methane under water deficient conditions. These materials display different methane reforming activities, and remain free of praseodymium, samarium and gadolinium oxide phases respectively after use in a reaction gas stream composed of 50% CH4–5% H2O – (in the absence and presence of 50 or 200 ppm H2S) – balance He at 740 °C. The results show that Ce0.8Pr0.2O2  δ, Ce0.85Sm0.15O2  δ and Ce0.9Gd0.1O2  δ are effective catalysts for reforming of methane and H2S in the feed promotes the catalytic activity. Ce0.8Pr0.2O2  δ appeared to attain the highest activity for methane reforming, a feature that is associated with the ability of praseodymium to undergo a red–ox (Pr4 +/Pr3 +) and spreading action in the cerium oxide host structure, possibly resulting in a red–ox relationship between the components.  相似文献   

12.
Fine powders (particle size of 100–200 nm) of BaYxZr1  xO3  x/2 (x = 0, 0.08, 0.16) were produced by solid-state reaction at 1000–1050 °C using nanocrystalline ZrO2 and BaCO3 raw materials. The powders were densified by means of the spark plasma sintering process resulting in dense and homogeneous submicron microstructures. Near full density ceramics with grain size < 300 nm were obtained by sintering at 1600 °C for 1–5 min.  相似文献   

13.
Density, surface tension and refractive index of the binary mixture of catalytic deactivated compounds with 1-ethyl-3-methylimidazolium acetate {[EMIM][OAc]} ionic liquid were measured at temperature of 298.15–323.15 K from which the derived thermodynamic properties including excess molar volume and deviation of surface tension and refractive index were calculated. The derived thermodynamic properties could be explained well by the interaction between similar and dissimilar aromatic structure of the molecules over the entire mole fraction of ILs. It was observed that all the catalytic deactivated compounds and water molecules have significant structural interaction with [EMIM][OAc] via CH?π bond interaction, π?π stacking and n?π interaction over the entire mole fraction of IL at T = 298.15 K. Further the composition of ionic liquid have significant influence on the interaction between dissimilar aromatic structure of molecules like pyridine, indoline and quinoline in liquid phase as compared to temperature. The surface tension increases in the order of: hiophene > pyridine > quinoline > pyrrole > indoline > water; while the refractive index increases in the order: pyridine < water < pyrrole < thiophene < indoline < quinoline. The deviation of surface tension was found to be inversely proportional to the deviation of refractive index at T = 298.15 K. From these results it was concluded that the structure of the ionic liquids is very important for extraction processes on catalytic deactivated compounds, especially for pyridine, indoline and quinoline as compared to water molecules.  相似文献   

14.
The ternary liquid–liquid equilibria (LLE) of the following systems were analytically determined at 298.15 K and atmospheric pressure using stirred and thermo-regulated cells: {dimethyl carbonate (DMC) + methanol + water}, {DMC + ethanol + water}, {DMC + 1-propanol + water}, {DMC + 2-propanol + water}, {DMC + 1-butanol + water} and {DMC + 2-butanol + water}. The experimental ternary LLE data were correlated with the non-random two liquid (NRTL) and UNIversal QUAsiChemical (UNIQUAC) activity coefficient models. In addition, the Bachman–Brown correlation was used to ascertain the reliability of the experimental data for each system.  相似文献   

15.
We have conducted experiments to obtain cloud-point data of binary and ternary mixtures for poly(isobornyl acrylate) [P(IBnA)] (Mw = 100,000) + isobornyl acrylate(IBnA) in supercritical carbon dioxide (CO2), P(IBnA) (Mw = 100,000) + dimethyl ether (DME) in CO2, P(IBnA) (Mw = 100,000) in propane and butane, and P(IBnA) (Mw = 1,000,000) in propane, propylene, butane and 1-butene at high pressure conditions. Phase behaviors for these systems were measured at a temperature range from 323.4 K to 474.1 K and pressure up to 296.7 MPa. The cloud-point curves of P(IBnA) (Mw = 100,000) + IBnA and DME in CO2 change from upper critical solution temperature (UCST) behavior to lower critical solution temperature (LCST) behavior as IBnA and DME concentration increases, and liquid–liquid–vapor phase behavior appears for the P(IBnA) (Mw = 100,000) + CO2 + 80.3 wt.% IBnA system. Phase behaviors of P(IBnA) and 50 wt.% IBnA in CO2 and P(IBnA) in propane and butane show the pressure difference in accordance with Mw = 1,000,000 and Mw = 100,000 of P(IBnA). Also, the solubility curves for IBnA in supercritical CO2 were measured at a temperature range of (313.2–393.2) K and pressure up to 22.86 MPa. The experimental results were modeled with the Peng–Robinson equation of state (PR-EOS) using a mixing rule including two adjustable parameters. The critical property of IBnA is estimated with the Joback–Lyderson method.  相似文献   

16.
Boron was doped into diamond films which were synthesized homoepitaxially on polished (100) diamond substrates by means of microwave plasma-assisted chemical vapor deposition (MPCVD) using trimethylboron as the dopant at a constant substrate temperature of 1073 K. The morphologies and electrical properties of the synthesized diamond films were dependent on the total reaction pressure. A maximum Hall mobility, 760 cm2 V−1 s−1, was obtained for the film synthesized at 10.7 kPa. The values of Hall mobility were comparable with those obtained for B2H6-doped films at corresponding hole concentrations.  相似文献   

17.
We newly synthesized a metal–organic framework (MOF) Rb2(adp)[Zn2(ox)3]·3H2O (adp = adipic acid; ox2  = oxalate), where the rubidium ions, carboxylic acid groups, and water molecules are located in an interlayer space of a two-dimensional (2-D) oxalate-bridged network. The structure of this compound was determined using single-crystal X-ray diffraction analysis. Hydrated phases of this compound were examined using thermogravimetry and water vapor adsorption measurements. Proton conductivity in this MOF was investigated by alternating current impedance measurements. Systematic comparison with previously reported isomorphous 2-D compounds A2(adp)[Zn2(ox)3]·3H2O (A = NH4 and K) showed that the difference in the ionic radii of the cations leads to a difference in activation energy of proton conductivity and that absence of NH4+ ions causes a significant decrease in proton conductivity, even though the ionic radius of Rb+ (1.52 Å) is closer to that of NH4+ (1.61 Å) than that of K+ (1.38 Å).  相似文献   

18.
The photoassisted degradation (HPLC-UV absorption), dehalogenation (HPLC-IC) and mineralization (TOC decay) of the flame retardants tetrabromobisphenol-A (TBBPA) and tetrachlorobisphenol-A (TCBPA) were examined in UV-irradiated alkaline aqueous TiO2 dispersions (pH 12), and for comparison the parent bisphenol-A (BPA, an endocrine disruptor) in pH 4–12 aqueous media to assess which factor impact most on the photodegradative process. Complete degradation (2.7–2.8 × 10−2 min−1) and dehalogenation (1.8 × 10−2 min−1) of TBBPA and TCBPA occurred within 2 h of UV irradiation, whereas only 45–60% mineralization (2.3–2.7 × 10−3 min−1) was complete within 5 h for the flame retardants at pH 12 and ca. 80% for the parent BPA. Factors examined in the pH range 4–12 that impact the degradation of BPA were the point of zero charge of TiO2 particles (pHpzc; electrophoretic method), particle or aggregate sizes of TiO2 (light scattering), and the relative number of OH radicals (as DMPO–OH adducts; ESR spectroscopy) produced in the UV-irradiated dispersion. Dynamics of BPA degradation (2.0–2.4 × 10−2 min−1) were pH-independent and independent of particle/aggregate size, but did correlate with the number of OH radicals, at least at pHs 4 to 8–9, after which the rates decreased somewhat at pH > 9 with decreasing adsorption owing to Coulombic repulsive forces between the very negative TiO2 surface and the anionic forms of BPA (pKas ca. 9.6–11.3), even though the number of OH radicals continued to increase at the higher pHs.  相似文献   

19.
Barium dititanate (BaTi2O5) thick films were prepared on a Pt-coated Si substrate by laser chemical vapor deposition, and ac electric responses of (0 2 0)-oriented BaTi2O5 films were investigated using several equivalent electric circuit models. BaTi2O5 films in a single phase were obtained at a Ti/Ba molar ratio (mTi/Ba) of 1.72–1.74 and deposition temperature (Tdep) of 908–1065 K as well as mTi/Ba = 1.95 and Tdep = 914–953 K. (0 2 0)-oriented BaTi2O5 films were obtained at mTi/Ba = 1.72–1.74 and Tdep = 989–1051 K. BaTi2O5 films had columnar grains, and the deposition rate reached 93 μm h?1. The maximum relative permittivity of the (0 2 0)-oriented BaTi2O5 film prepared at Tdep = 989 K was 653 at 759 K. The model of an equivalent circuit involving a parallel combination of a resistor, a capacitor, and a constant phase element well fitted the frequency dependence of the interrelated ac electrical responses of the impedance, electric modulus, and admittance of (0 2 0)-oriented BaTi2O5 films.  相似文献   

20.
The electrical properties and oxygen permeability of glass–ceramics 55SiO2–27BaO–18MgO, 55SiO2–27BaO–18ZnO and 50SiO2–30BaO–20ZnO (%mol), which possess thermal expansion compatible with that of yttria-stabilized zirconia (YSZ) solid electrolytes, were studied between 600 and 950 °C in various atmospheres. The ion transference numbers, determined by the modified electromotive force (e.m.f.) technique under oxygen partial pressure gradients of 21 kPa/(1–8) × 102 Pa and 21 kPa/(1 × 10−18–2 × 10−12) Pa, are close to unity both under oxidizing and reducing conditions. The electronic contribution to the total conductivity increases slightly on increasing temperature, but is lower than 2% and 7% for the Zn- and Mg-containing compositions, respectively. The conductivity values measured by impedance spectroscopy vary in the range (1.4–7.8) × 10−6 S/cm at 950 °C under both oxidizing and reducing conditions, with activation energies of 122–154 kJ/mol and a minor increase in H2-containing atmospheres, indicating possible proton intercalation. In agreement with the electrical measurements which indicate rather insulating properties of the glass–ceramics, the oxygen permeation fluxes through sintered sealants and through sealed YSZ/glass–ceramics/YSZ cells are very low, in spite of an increase of 15–40% during 200–230 h under a gradient of air/H2–H2O–N2 due to slow microstructural changes.  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