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 15 毫秒
1.
The ionic conductivity of PAN‐TiO2‐LiClO4 as a function of TiO2 concentration and temperature has been reported. The electrolyte samples were prepared by solution casting technique. Their conductivity was measured using the impedance spectroscopy technique. The highest room temperature conductivity of 1.8 × 10?4 S cm?1 was obtained at 7.5 wt % of TiO2 filler. It was observed that the relationship between temperature and conductivity were linear, fitting well in Arrhenius and not in Vogel‐Tamman‐Fulcher equation. The pre‐exponential factor, σ0 and Ea are 1.8 × 10?4 S cm?1 and 0.15 eV, respectively. The conductivity data have been supported by differential scanning calorimeter (DSC) analysis. DSC analysis showed that there was a significant change in glass transition temperature (Tg) with the filler concentration. The SEM micrograph revealed that the TiO2 particles are dispersed in the electrolyte, thus enhancing its conductivity. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

2.
An electron paramagnetic resonance (EPR) study of the photoexcited triplet state of four free base porphyrins is presented. The zero field splitting parameters (ZFS) |D| and |E| were calculated from the EPR spectra of the porphyrins dissolved in n-octane matrices at 80°K. |D| = 0.0359 cm?1, |E| = 0.0079 cm?1 for tetra phenyl porphyrin (H2 TPP), |D| = 0.0432 cm?1, |E| = 0.0037 cm?1 for tetra (per-fluoro) phenyl porphyrin H2T (per-F) PP, |D| = 0.0366 cm?1, |E| = 0.0078 cm?1 for tetra (para-chloro) phenyl porphyrin H2T(P-Cl)PP, |D| = 0.0369 cm?1, |E| = 0.0076 cm?1 for tetra (para-methyl) phenyl porphyrin H2T(P-Me)PP. The transient behavior of the EPR signal intensities in the last two porphyrins is discussed. The depopulation rate constants of the triplet sublevels kp, the ratio between the population rate constants Ap (at zero field, p = x,y,z), and the spin lattice relaxation rate W within the triplet manifold, were calculated. kx = (12 ± 2) × 102 sec?1, ky = (0.5 ± 0.1) × 102 sec?1, kz = (1.2 ± 0.4) × 102 sec?1, Ax:Ay:Az ? 0.63:0.01:0.33, W = (0.4 ± 0.1) × 104 sec?1 for H2T(P-Cl)PP, kx = (7 ± 2) × 102 sec?1, ky = (4 ± 1) × 102 sec?1, kz = (1.5 ± 0.5) × 102 sec?1, Ax:Ay:Az ? 0.56:0.31:0.13, W = (1.7 ± 0.4) × 103 sec?1 for H2T(P-Me)PP.  相似文献   

3.
An aerochitin–titania (TiO2) composite was successfully synthesized and characterized by Fourier transform infrared spectroscopy, X‐ray diffraction, thermogravimetric analysis, field emission scanning electron microscopy, and N2 adsorption isotherms. The photocatalytic activity of the composite was investigated on the degradation of the model organic pollutant, methylene blue (MB) dye, under UV irradiation. The aerochitin–TiO2 composite showed excellent adsorptive and photocatalytic activity with a degradation degree of 98% for MB. The first‐order rate constants for the photodegradation MB by TiO2 nanoparticles and aerochitin–TiO2 composite were found to be (3.49 ± 0.04) × 10?3 and (1.82 ± 0.02) × 10?2 min?1. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2018 , 135, 45908.  相似文献   

4.
BACKGROUND: Thermodynamics and kinetics data are both important to explain the extraction property. In order to develop a novel separation technology superior to current extraction systems, many promising extractants have been developed including calixarene carboxylic acids. The extraction thermodynamics behavior of calix[4]arene carboxylic acids has been reported extensively. In this study, the mass transfer kinetics of neodymium(III) and the interfacial behavior of calix[4]arene carboxylic acid were investigated. RESULTS: The rate constant (Kao) becomes constant when the stirring speed was controlled between 250 rpm and 400 rpm. The activation energy (Ea) was calculated to be 21·41 kJ mol?1 or 88·17 kJ mol?1 (dependent on temperature) from the slope of log Kao against 1000/T. The linear relationship between the specific area and the extraction rate is the characteristic of an interfacial reaction control. The minimum bulk concentration of the extractant necessary to saturate the interface (Cmin) is lower than 4·19 × 10?4 mol L?1. CONCLUSION: The effect of stirring speed, temperature, and species concentration on the extraction rate demonstrates that the extraction regime depends on the extraction conditions. The chemical reaction control governs the extraction regime at temperatures below 303 K and a mixed control regime occurs when the temperature is between 303 K and 318 K. The probable locale for the chemical reaction is at the liquid–liquid interface and the rate equation is deduced to be: ? d[Nd3+](a)/dt = kf[Nd3+](a)[H4A](o)0·727[H+](a)?0·978. The rate‐controlling step was suggested by the analysis of the experimental results. Copyright © 2008 Society of Chemical Industry  相似文献   

5.
Continuous crystallization of potassium carbonate was investigated in a mixed suspension mixed product removal (MSMPR) crystallizer with 7.0 1 working volume. Highly polydispersed product was obtained, with x 50 varying between 320 and 670 μm. Crystal growth rates can be described by the model of Abegg, Stevens and Larson (ASL) (size dependent crystal growth rate). Values of G0 = 8.6 × 10?9 to 3.6 × 10?8 m/s and B0/φ = 2.7 × 108 to 4.2 × 109 1/m3s were obtained for mT = 45 to 92 kg/m3 and ε = 0.45 ± 0.05 W/kg. Nucleation kinetics at 27 °C can be described by the equation: B0/φ = kBφsε0.73G02.5. Since G0 ∝ σ, convection and/or diffusion rather than surface integration are the crystal growth controlling mechanisms.  相似文献   

6.
《分离科学与技术》2012,47(18):2531-2541
Abstract

Kinetics of silver extraction from nitrate solutions with triisobutylphosphine sulfide dissolved in n-octane was studied. Experiments were carried out in a rotating diffusion cell. The rate constants of forward and reverse chemical reactions were evaluated: the former k? f = 1.064 × 10?3 m9/(mol3·s) and the latter k? f = 2.085 × 10?1 s?1. The value obtained for the activation energy shows that the process of silver extraction with triisobutylphosphine sulfide is a predominantly diffusion-controlled process.  相似文献   

7.
A preparative size‐exclusion chromatography (SEC) column (500 × 10 mm) packed with regenerated cellulose gel particles was used for the separation of superoxide dismutase (SOD) from pig blood. The fraction F‐5, with an Mw of 12.4 × 104, composed of more aspartic acid (Asp), glutamine (Glu), and tyrosine (Tyr) but less cysteine (Cys), has the highest specific activity (204 units per mg protein), which was higher than that of the commercial product (175 units per mg protein). The change of fluorescence intensities at zero concentration of the fractions was identical to that of the activities. The results from amino acid sequence analysis and size‐exclusion chromatography combined with static laser light scattering (SEC–LLS) analysis indicated that the separation mechanism of SOD is based on both size exclusion and ion affinity. They also showed that the activity of SOD is determined mainly by its composition. At a flow rate of 0.62 mL min?1, 42 mg of a high specific activity product can be obtained from crude SOD in 1 day. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 763–768, 2003  相似文献   

8.
The thermal expansion of La0.5Sr0.5Co0.25Fe0.75O3 (LSCF55) is investigated both by first principles phonon calculations combined with the quasi‐harmonic approximation (QHA) and by experimental approaches. Within the framework of the QHA, the volumetric thermal expansion coefficient of rhombohedral LSCF55 is calculated as αV,GGA = 50.34 × 10?6 K?1. For comparison, the lattice expansion and the volume expansion of LSCF55 grain are measured by in situ high‐temperature X‐ray diffractometer (HT‐XRD). An anisotropic thermal expansion of rhombohedral LSCF55 with αa,hex = 10.89 × 10?6 K?1 and αc,hex = 21.18 × 10?6 K?1 is obtained. The volumetric thermal expansion coefficient is measured as αV,HT‐XRD = 43.17 × 10?6 K?1. In addition, the effectively isotropic expansion coefficients of a polycrystalline LSCF55 bar specimen are measured using a vertical high‐performance thermo‐mechanical analyzer and yield αl,bar specimen = 17.37 × 10?6 K?1 and αV,bar specimen = 52.11 × 10?6 K?1.  相似文献   

9.
A comparative study was carried out in order to evaluate the kinetics of the formation of a number of primary and secondary oxidation products during oxidation of olive oil in the Rancimat test at 100–130 °C. There were good correlations between the Rancimat index (OSI) and stability indices (IP) measured in the Rancimat test with no significant differences in kinetic parameters calculated from them. Mean values of the temperature coefficient, Q10 number, activation energy (Ea), frequency factor (A), and free energy of activation (ΔG++) for olive oil oxidation were calculated to be ?3.44 × 10?2°C?1, 2.21, 98.91 kJ/mol, 12.17 × 1012 h?1, and 128.25 kJ/mol, respectively. Each unit change in Ea was accompanied by an average 1.43 × 1012 change in A, indicating a higher contribution for factor A than for Ea to the olive oil stability. The Ea and A correlated well with the values of enthalpy and entropy, respectively. The values of OSI or IP could be described well by the ΔG++ values. Kinetic data indicated that olive oil stability is more affected by the indigenous antioxidants than by the fatty acid composition.  相似文献   

10.
Static and dynamic light‐scattering techniques were used to study biodegradable thermoplastic poly(hydroxy ester ether) in N,N‐dimethylacetamide (DMAc). A weight‐average molecular weight MW = 6.4 × 104 g/mol, radius of gyration RG = 9.4 nm, second‐virial coefficient A2 = 1.05 × 10?3 mol mL/g2, translational diffusion coefficient D = 1.34 × 10?7 cm2/s, and hydrodynamic radius RH = 8.3 nm are reported. In addition, the effect of H2O on the polymer chain's conformation and architecture in a DMAc/H2O solution is evaluated. Results suggest that H2O makes the mixed solvent poorer as well as promotes polymer chain branching via intramolecular transesterification. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 80: 1737–1745, 2001  相似文献   

11.
The kinetic of the reaction of sand with aqueous NaOH corresponding to the ratio SiO2/Na2O = 2 was studied in a pressure vessel at 220°C and 2.7 MPa. Since the kinetic curves could not be obtained directly from the experimental data, a new method is proposed to plot the entire kinetic graph from experimental data. An analytical expression of the type α = A [1 -exp(-Bt)] describes the system perfectly. The constants A and B were calculated for a silica sample having a narrow granulometric distribution (range i.e. 300–315 μm). The value of A is found to be almost constant, between 0.95 to 0.99 and B ranges from 0.03 to 0.14 when [OH-] increases from 0.5 to 12.5 mol/L. The kinetic order with respect to OH? is equal to 0.470 + 0.013 and the kinetic constant at 220°C is 3.933 × 10?6 g/m2.s.  相似文献   

12.
A combination of polarized Raman technique, infrared reflectance spectra, and first‐principles density‐functional theoretical calculations were used to investigate structure transformation and lattice vibrations of Na0.5Bi0.5TiO3, Na0.5Bi0.5TiO3–5%BaTiO3, and Na0.5Bi0.5TiO3–8%K0.5Bi0.5TiO3 single crystals. It was found that Na0.5Bi0.5TiO3 is of a two‐phase mixture with rhombohedral and monoclinic structures at room temperature. Correspondingly, three Raman‐active phonon modes located at 395, 790, and 868 cm?1, which were previously assumed as A1 modes of rhombohedral phase have been reassigned as A′′, A, and A modes of monoclinic phase in the present work. In particular, a strong low‐frequency A′′ mode at 49 cm?1 was found and its temperature dependence was revealed. Two deviations from linearity for the abrupt frequency variation in the A′′ mode and Ti–O bond have been detected at temperatures of ferroelectric to antiferroelectric phase transition TF–AF and dielectric maximum temperature Tmax. The appearance of Na–O vibrations at 150 cm?1 was found below Tmax, indicating the existence of nanosized Na+TiO3 clusters. The observed Raman and infrared active modes belonging to distinct irreducible representations are in good agreement with group‐theory predictions, which suggests 9A1+9E and 36A′′+24A modes for the rhombohedral and monoclinic phases of Na0.5Bi0.5TiO3, respectively.  相似文献   

13.
Different molecular weight phenolphthalein poly(ether ketone) (PEK-C) and phenolphthalein poly(ether sulfone) (PES-C) fractions in chloroform (CHCl3) were studied by static and dynamic laser light scattering (LLS). The dynamic LLS revealed that both the PEK-C and PES-C samples contain some large polymer clusters formed in the process of polymerization. These large clusters can be removed by filtering the solution with a 0.1-μm filter. The positive second virial coefficient (A2) shows that CHCl3 is a good solvent for these polymers at room temperature. The persistence length and the Flory characteristic ratio of these polymers in CHCl3 at 25°C are ~2 nm and ~3, respectively, which indicate that these polymer chains are flexible. A combination of static and dynamic LLS results, namely the weight-average molecular weight from static LLS and the line-width distribution from dynamic LLS, lead to two calibrations between the translational diffusion coefficient (D) and molecular weight (M): D = 2.20 × 10?4 M?0.56 for PEK-C and D = 2.45 × 10?4 M?0.55 for PES-C in CHCl3. Using these calibrations, we are able to estimate not only the molecular weight distributions of these fractionated polymers, but also unfractionated PEK-C and PES-C samples.  相似文献   

14.
The liquid–liquid extraction of Gd(III) from aqueous nitrate medium was studied using di‐2‐ethylhexylphosphoric acid (HDEHP) in kerosene. On the basis of the slope analysis data, the composition of the extracted species was found to be [Gd A3(HA)] with the extraction equilibrium constant (Kex) = (1.48 ± 0.042) × 10?12 mol dm?3. The results of the effect of temperature on the value of the equilibrium extraction constant indicated the endothermic character of the extraction system. The kinetics of the forward extraction of Gd3+ from nitrate medium by HDEHP in kerosene was investigated using the single drop column technique. The rate of flux (mass transfer per unit area) was found to be proportional to [Gd(III)], [H2A2](o), [NO3?], and [H+]?1 in the liquid drop organic phase. The forward extraction rate constant, kf, was 2.24 × 10?3 m s?1 using the equation: Copyright © 2005 Society of Chemical Industry  相似文献   

15.
High‐resolution synchrotron powder X‐ray diffraction (XRD) experiments were conducted to clarify the transformation of sillimanite to mullite (mullitization) and determine the mullitization temperature (Tc). We were able to distinguish sillimanite and mullite in the XRD patterns, despite their very similar crystallographic parameters, and to detect the appearance of small mullite peaks among sillimanite peaks. Analysis of the Johnson‐Mehl‐Avrami (JMA) equation for mullitization ratio (ζ) revealed that at temperatures T≥1240°C the mullitization had the same kinetics. The activation energy E at T≥1240°C obtained from the Arrhenius plot was 679.8 kJ mol?1. In analysis using a time‐temperature‐transformation diagram for mullitization, a mullitization curve of ζ=1% can be described as where t is time, n is a reaction‐mechanism‐dependent parameter determined as 0.324 by JMA‐analysis, k0 is the frequency factor, EA is the activation energy for atomic diffusion, and represents the activation energy for nucleation. The results of fitting the data to this equation were Tc=1199°C, A=3.9×106 kJ mol?1 K?2, EA=605 kJ mol?1, and k0=3.65×1015. We conclude that the boundary between sillimanite and mullite+SiO2 in the phase diagram is ~1200°C.  相似文献   

16.
Constant strain-rate uniaxial extension tests to rupture were performed at 23°C on a plasticized poly(vinyl chloride) (PVC) in the glass-to-rubber transition zone, Tg = ?18°C, where experimental failure time tf is equal to or greater than material relaxation time τ. Range of strain rate is from 1.8 × 10?4 to 1.8 × 10?1 sec?1. The rupture characteristics in stress/strain time space are analyzed on the three coordinate planes. Time and deformation separability are examined in a nonlinear-constitutive relation. Rupture data on the coordinate planes are described by the Bueche-Halpin theory in which time effects are considered through a small-deformation viscoelastic property. Fracture surface morphology and separation processes in the crack tip are related to rupture characteristics.  相似文献   

17.
Molar conductances have been determined for solutions of Aerosol-OT in n-propanol at 25°C. Analysis of these results with the Fuoss—Hsia conductance equation using a distance parameter d = 13.70 × 10?10 m yields a limiting molar conductance Λ = 17.5 ± 0.5 Ω?1 and an association constant KA = (27 ± 3) × 102 dm3 mol?1. The latter is substantially larger than that calculated with the Bjerrum equation.  相似文献   

18.
This study investigated the adhesion between zirconia framework and four veneering ceramic (VC) materials with varying coefficients of thermal expansions (CTE). Zirconia rods (N?=?40) (ICE Zirkon) (diameter: 4 mm, height: 20 mm) were milled and sintered. After firing, the zirconia rods were air-abraded and cleaned. They were randomly assigned to receive four VCs (n?=?10/group), namely (a) Vita VM9 (VZ; 9–9.2?×?10?6? K?1), (b) Cerabien ZR (CZ; 9.1?×?10?6 K?1), (c) Matchmaker ZR (MM; 9.4?×?10?6?K?1), and (d) Ice Zirconia Ceramic (IZ; 9.6?×?10?6?K?1). The VCs were then fired onto zirconia rods (height: 2 mm, thickness: 2 mm) circumferentially and were thermocycled for 6000 times (5/55 °C, dwell time: 30?s). Specimens were loaded from the top of the zirconia rods (0.5 mm/min) in a universal testing machine until debonding. Shell–Nielsen bond strength values were calculated (MPa). Failure types were evaluated under SEM. The data were statistically analyzed (one-way ANOVA, Tukey’s; α?=?0.05). Weibull distribution values including the Weibull modulus (m) (0.05) was calculated. The highest mean bond strength (MPa) was obtained for CZ (42.08?±?4.08), followed by VZ (41.77?±?4.92), MM (40.7?±?3.64), and IZ (40.05?±?5.78). While mean bond strength for VZ, MM, and IZ were not significantly different (p?>?0.05), CZ was significantly higher than that of IZ (p?<?0.05). The lowest shape value was for VZ (m?=?16.94) and the highest for MM (m?=?20.16). Mainly, adhesive failures followed by mixed failures were observed. VCs with a greater mismatch of CTE with the zirconia framework exhibited similar Shell–Nielsen bond strength to those with fewer mismatches. CTE mismatch did not affect the results of CZ (9.1?×?10?6 K?1) and IZ (9.6?×?10?6 K?1).  相似文献   

19.
The graft copolymer of xanthan gum with methacrylic acid was synthesized in inert atmosphere by using Fentos reagent as a redox initiator. The effect of reaction conditions on grafting parameters [G(%), E(%), C(%), A(%), H(%), and Rg] was investigated. Similar trend was observed on increasing the concentration of ferrous ion and hydrogen peroxide from 4.0 to 20.0 × 10?3 mol dm?3 and 2.5 to 10 × 10?3 mol dm?3 respectively, i.e., initially grafting parameters increased and after a certain range of concentration grafting parameters showed decreasing trend. Hydrogen ion shows influenced result i.e., small increment of concentration in hydrogen ion presents much increment in percent of grafting. It was observed that the [G(%), E(%), C(%), A(%), and Rg] increased upto 6.67 × 10?2 mol dm?3 concentration of methacylic acid after that it decreased. Maximum G(%) was obtained at minimum concentration of xanthan gum i.e., at 40 × 10?2 g dm?3. The optimum temperature and time duration of reaction for maximum percentage of grafting were found to be 45°C and 150 min respectively. Thermogravimetric analysis showed that the xanthan gum‐g‐methacrylic acid is thermally more stable than pure gum. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2007  相似文献   

20.
Experiments at various Sb2O3 concentrations were made in a pilot plant and the effect of Sb2O3 on continuous esterification between terephthalic acid (TPA) and ethylene glycol (EG) was obtained. Reaction rate constants of the previously reported reaction scheme were determined to fit with the experimental data obtained. It was found that the effect of Sb2O3 on reaction rate constant (ki) can be expressed as follows.
  • k1 = (3.75 × 10?4Sb + 1.0) × 1.5657 × 109exp(?19,640/RT)
  • k2 = (4.75 × 10?4Sb + 1.0) × 1.5515 × 108exp(?18,140/RT)
  • k3 = (6.25 × 10?4Sb + 1.0) × 3.5165 × 109exp(?22,310/RT)
  • k4 = (4.50 × 10?4Sb + 1.0) × 6.7640 × 107exp(?18,380/RT)
  • k5 = (3.50 × 10?4Sb + 1.0) × 7.7069 × exp(?2810/RT)
  • k6 = (1.75 × 10?4Sb + 1.0) × 6.2595 × 106exp(?14.960/RT)
  • k7 = (3.75 × 10?4Sb + 1.0) × 2.0583 × 1015exp(?42,520/RT)
Simulation of esterification with these reaction rate constants at various Sb2O3 concentrations was made and the following results were obtained.
  • 1 Sb2O3 accelerates the esterification reaction between TPA and EG.
  • 2 Sb2O3 accelerates the main reaction and its effects on side reactions are minor. The higher the addition rate of Sb2O3, the lower the carboxyl end-group concentration (AV) and diethylene glycol content (DEG).
  • 3 The comparison between simulation with potassium titanium oxyoxalate (PTO) in the previous report and with Sb2O3 in the present report shows that the acceleration of polycondensation reaction by Sb2O3 is higher. DEG formation rate is lower with PTO than Sb2O3.
  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