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.
Thermal oxidative degradation behavior of polypropylene (PP) with different tacticities was studied based on the activation energy (ΔE) data obtained by thermogravimetric analysis (TGA). The ΔE value showed a negative proportion to the content of meso pentad fraction (mmmm) in all of isotactic PP (iPP) samples, and that of syndiotactic PP sample considerably deviated from this negative proportion relationship. Since the value of mmmm was directly related to polymer chain conformation, the ΔE value was thought to have close connection with the concentration of 31 helix conformation in the iPPs. The ΔE changes would be caused by the competition between uni- and bimolecular hydroperoxide decomposition, which was controlled by concentration and character of conformations of PPs.  相似文献   

2.
The electrosorption properties of p-norborn-2-yl phenolate ions in alkaline solutions were investigated by ac polarographic and electrocapillary measurements.

Two adsorption regions were found. At low bulk surfactant concentrations the adsorption at the positively charged electrode (−0.2 E −0.6 V) is predominant while at higher surfactant concentrations the adsorption at the negatively charged electrode (−0.6 E −1.0 V) is more pronounced. At E = −0.40 V the adsorption parameters were determined (a ≈ 2; ΔG°A = −32.5 ± 1 kJ mol−1. Between −0.6 E −1.0 V one potential of maximum adsorption for all concentrations does not exist and therefore the adsorption parameters could not be calculated.

At E = −0.40 V progressive two-dimensional nucleation with a nucleation order of 3 was observed which corresponds well with the high attraction constant.

The electrode reaction S2O2−8 + 2e → 2 SO2−4 is inhibited by norborn-2-yl phenolate ions in the potential range −0.2 E −0.6 V. In the second potential range of capacity decrease the electrode process is much less retarded. At E = −0.40 V, in a similar manner as described for neutral molecules, a linear dependence of the log ks (ks apparent rate constant) on ln cA and π (π = surface film pressure), respectively, has been found.  相似文献   


3.
It is shown that there is limited validitity to the doctrine that true interfacial separation, in an adhering system, is highly improbable. An analysis employing the Griffith-Irwin crack theory yields these results: The important parameters are, difference in elastic moduli, ΔE; differences in g, the energy dissipation per unit crack extension; thickness, Δ1 or δ2, of the region where dissipation occurs; and the presence or absence of strong interfacial bonds. If the forces across the interface are appreciably weaker than the cohesive forces in either phase, there is a strong minimum in g at the interface. For flaws of equal size, an interfacial flaw will be the site of initiation of failure. If strong interfacial bonds are present, then if Δg and ΔE have the same sign, failure is most probable, deep within one phase. If Δg and ΔE have opposite signs, failure may be initiated, and may propagate, at a distance δ from the interface, in the phase with lower g. This may be mistaken for weak-boundary layer failure.  相似文献   

4.
Light scattering by monodisperse solutions of rigid rod-like anisotropic macromolecules, with linear dimensions l of the order of incident wavelength λ, oriented in an external d.c. electric field, →E has been analysed. The relative variations δVEv, δHEv, δVEh, δHEh of the scattered light components are discussed for the three values [l/λ] = 1, 0·5, 2, and various reorientation parameters, p = [μE/kT] of the permanent dipole moment μ and q = [(31)E2/2kT] of the moment induced by the principal polarizabilities 1 = 23. The saturation orientation field strength has been calculated for certain macromolecules with the aim of determining their optical anisotropy numerically.  相似文献   

5.
Deformation kinetics of ageing materials   总被引:1,自引:0,他引:1  
K.C. Valanis  S.T.J. Peng 《Polymer》1983,24(12):1551-1557
A constitutive equation of time-dependent, chemically stable materials, which stems from the basic ideas of the irreversible thermodynamics of an internal variable and Eyring's absolute reaction rate theory, has been extended to chemically unstable materials. This formulation is quite general and, in principle, can be applied to many types of materials. In this paper, the ageing behaviour of time-dependent network polymers undergoing chain scission is considered. In the network scission process, we postulate that the energy barrier is affected by a changing of the chemical crosslink density. An explicit equation to account for the energy barrier change, which influences the relaxation process, is formulated. For the purpose of illustration, the effect of different chemical crosslink density, ν, on the relaxation rate has been considered, from which the following theoretical expression of relaxation modulus ΔE(t) is obtained:

ΔE = ΔE[t expv/kT)]

It can be seen that a change in v leads to an effective change in the time scale, usually denoted by ax. Here the analytical expression ax = exp(γν/kT) correlates quite well with the experimental data.  相似文献   


6.
A Monte Carlo computer simulation model for the electrophoretic deposition of polymer chains on a discrete lattice is used to study the polymer density profile, interface growth, and its dependence on field, temperature, and molecular weight. The interface width (W) decreases WE−1/2 on increasing the field (E). Width (W) depends non-monotonically on the temperature (T): a power-law decay is followed by a power-law increase on raising the temperature. Monotonic decay of the interface width with the molecular weight is possibly a stretched exponential. Conformation and dynamics of a tracer chain is used to probe its characteristics in interface to bulk region. The root mean square (rms) displacement of the center of mass of the tracer chain shows an ultra-slow motion, Rtν (ν0.1–0.01 at E=0.1–1.0) as the driven chain moves deeper from interface to bulk. Longitudinal compression of the radius of gyration (Rg) of the chain increases with the field; transverse components (Rgx, Rgy) are larger than the longitudinal component (Rgz). The transverse component (Rgx(y)) becomes oscillatory due to periodic squeezing at high fields as the field competes with the polymer barriers.  相似文献   

7.
Water–gas shift reaction was studied over two nanostructured CuxCe1−xO2−y catalysts: a Cu0.1Ce0.9O2−y catalyst prepared by a sol–gel method and a Cu0.2Ce0.8O2−y catalyst prepared by co-precipitation method. A commercial low temperature water–gas shift CuO–ZnO–Al2O3 catalyst was used as reference. The kinetics was studied in a plug flow micro reactor at an atmospheric pressure in the temperature interval between 298 and 673 K at two different space velocities: 5.000 and 30.000 h−1, respectively. Experimentally estimated activation energy, Eaf, of the forward water–gas shift reaction at CO/H2O = 1/3 was 51 kJ/mol over the Cu0.1Ce0.9O2−y, 34 kJ/mol over the Cu0.2Ce0.8O2−y and 47 kJ/mol over the CuO–ZnO–Al2O3 catalyst. A simple rate expression approximating the water–gas shift process as a single reversible surface reaction was used to fit the experimental data in order to evaluate the rate constants of the forward and backward reactions and of the activation energy for the backward reaction.  相似文献   

8.
Electrical resistivity and Seebeck (S) measurements were performed on (La1−xSrx)MnO3 (0.02x0.50) and (La1−xSrx)CoO3 (0x0.15) in air up to 1073 K. (La1−xSrx)MnO3 (x0.35) showed a metal-to-semiconductor transition; the transition temperature almost linearly increased from 250 to 390 K with increasing Sr content. The semiconductor phase above the transition temperature showed negative values of S. (La1−xSrx)CoO3 (0x0.10) showed a semiconductor-to-metal transition at 500 K. Dominant carriers were holes for the samples of x0.02 above room temperature. LaCoO3 showed large negative values of S below ca. 400 K, indicative of the electron conduction in the semiconductor phase.  相似文献   

9.
Preliminary studies on a series of nanocomposite BaO–Fe ZSM-5 materials have been carried out to determine the feasibility of combining NOx trapping and SCR-NH3 reactions to develop a system that might be applicable to reducing NOx emissions from diesel-powered vehicles. The materials are analysed for SCR-NH3 and SCR-urea reactivity, their NOx trapping and NH3 trapping capacities are probed using temperature programmed desorption (TPD) and the activities of the catalysts for promoting the NH3 ads + NO/O2 → N2 and NOx ads + NH3 → N2 reactions are studied using temperature programmed surface reaction (TPSR).  相似文献   

10.
A new method is proposed here for determining counter-ions permselectivity through an ion exchange membrane. The theoretical base of the method is Henderson equation. After simplifying at special experimental conditions, a linear relationship between the bi-ionic membrane potential term and concentration ratio, i.e., [exp(zAF(−ΔET)/RT)−1]∝[CA,L/CB,L] can be obtained, from the slope of which, the permselectivity of counter-ions can be achieved. This method is very simple in operation and less time and money consuming. The method is exemplified by common 1–1, 2–2 and 2–1 counter-ions pair through a novel and published anion exchange membrane. The accuracy and the validity of the method is tested by three aspects: (1) the experimental [exp(zAF(−ΔET)/RT)−1]∝[CA,L/CB,L] curve approaches a straight line (2) the predicted order of counter-ions conforms to the results evaluated in a practical electrodialysis process; and (3) the results calculate from different referenced counter-ions for the same counter-ions pair agree well with each other.  相似文献   

11.
The electronic states of LaMn1−xCuxO3+λ (x=0–0.4) have been studied with X-ray photoelectron spectroscopy (XPS). The valence states of substituted copper ions were Cu2+ and the manganese ions were a highly mixed state of Mn3+ and Mn4+. The nonstoichiometry and electronic state of lattice oxygen have been studied. The samples at x=0 and 0.1 had an excess of lattice oxygen but those at x=0.2–0.4 had lattice oxygen deficiency. A modified Auger parameter (Δ′) was used to evaluate the electronic states of oxygen ions. The Δ′ of lattice oxygen increased with increasing substitute quantity. This increase of Δ′ reflected the decrease of ionic bond character of lattice oxygen. The adsorbed oxygen species on LaMn1−xCuxO3+λ was assigned mainly as O from the peak positions of spectra for the O 1s and O KLL levels, and the Δ′ of this O decreased with x. This decrease, i.e., the increase of ionic bond character of adsorbed oxygen was correlated well with the value of nonstoichiometry of lattice oxygen.

The rate of CO oxidation at 448 K was increased by the substitution till x=0.4. We consider that this enhancement of reactivity comes from the change of electronic state of adsorbed oxygen, O itself, i.e., a weak interaction between O and low coordinated metal site brings about a high reactivity.  相似文献   


12.
Rate constants or exchange current densities of electrode surface processes involving adatom arrays are conveniently evaluated by determining that sweep rate, s0 (the reversibility parameter), in a linear potential sweep (LPS) experiment, below which the process just remains kinetically reversible, ie its overpotential is sensibly zero. Transition to irreversibility is characterized by peak potentials, Ep, becoming linear in the log of the sweep rate, s, following a region of independence of s for < s0. A suitable extrapolation procedure enables s0 to be evaluated. However, if s0 is large and/or the resistivity of the solution is appreciable, the IRu drop associated with uncompensated resistance in the measurement system can be comparable with the increase of Ep with log s, when s > s0, rendering evaluation of s0 inaccurate. While compensation or empirical correction for this may be made, it is desirable that the nature of the IRu effect of the LPS I vs E profiles be understood in a more fundamental way. It is the purpose of this communication to provide such a treatment of this effect through evaluation of the actual time-dependent potential that becomes applied to the electrode, and to propose criteria based on the product of s0, Ru and the reaction pseudocapacitance, Cφ for indicating the anticipated extent of the IRu effect in the evaluation of s0. While the transition in the Ep vs log s plot may be due to IRu effects as well as to kinetic irreversibility, when the former are appreciable, it is shown that the corresponding transition in the value of Cφ with increasing log s can usually allow spurious IRu and significant irreversibility effects to be distinguished.  相似文献   

13.
A range of hydroxy-terminated polybisphenol A terephthalate and isophthalate blocks have been prepared with molecular weights 800–5000, which were then coupled with phosgene to give alternating polyester copolycarbonates. These materials have been characterized by their physical, thermal and mechanical properties. The thermal properties have been investigated using differential scanning calorimetry and the glass transition temperatures, specific heats and Δcps′ (at the Tg) values obtained. A relationship appears to exist between Δcp and the reciprocal molecular weight of the polyester blocks, and the molar ratio of ‘ester’ to carbonate of the copolycarbonates. A maximum in the glass transition temperature has also been observed in the copolycarbonates, corresponding to a certain ‘ester’: CO3 ratio. Tensile mechanical analyses have been performed on cast or moulded films of the copolycarbonates. The terephthalates can give films which extend uniformly but the isophthalates always neck.  相似文献   

14.
The catalytic effect of a heteropolyacid, H4SiW12O40, on nitrobenzene (20 and 30 μM) oxidation in supercritical water was investigated. A capillary flow-through reactor was operated at varying temperatures (T=400–500 °C; P=30.7 MPa) and H4SiW12O40 concentrations (3.5–34.8 μM) in an attempt to establish global power-law rate expressions for homogenous H4SiW12O40-catalyzed and uncatalyzed supercritical water oxidation. Oxidation pathways and reaction mechanisms were further examined via primary oxidation product identification and the addition of various hydroxyl radical scavengers (2-propanol, acetone, acetone-d6, bromide and iodide) to the reaction medium. Under our experimental conditions, nitrobenzene degradation rates were significantly enhanced in the presence of H4SiW12O40. The major differences in temperature dependence observed between catalyzed and uncatalyzed nitrobenzene oxidation kinetics strongly suggest that the reaction path of H4SiW12O40-catalyzed supercritical water oxidation (average activation Ea=218 kJ/mol; k=0.015–0.806 s−1 energy for T=440–500 °C; Ea=134 kJ/mol for the temperature range T=470–490 °C) apparently differs from that of uncatalyzed supercritical water oxidation (Ea=212 kJ/mol; k=0.37–6.6 μM s−1). Similar primary oxidation products (i.e. phenol and 2-, 3-, and 4-nitrophenol) were identified for both treatment systems. H4SiW12O40-catalyzed homogenous nitrobenzene oxidation kinetics was not sensitive to the presence of OH√ scavengers.  相似文献   

15.
The ferrocene-ferricinium electrode (Pt/Foc, Fic+) was investigated in water, acetonitrile, ethanol, DMSO and DMF using single scan cyclic voltammetry and phase sensitive ac polarography. The oxidation-reduction is pseudo-reversible in all five solvents with an electrochemical rate constant of approximately 10−2 cm/s. In all solvents a slow irreversible chemical step involving the ferricinium cation follows electron transfer, so that slow cyclic voltammetry or polarography rather than potentiometry is preferred if ferrocene is to be used as a reference electrode in non-aqueous solvents.

The Strehlow assumption, ΔGtr(Foc) = ΔGtr(Fic+ gives very different free energies of transfer of single ions from non-aqueous solvents to water when compared with the TATB assumption that ΔGtr(Ph4As+) = ΔGtr(Ph4B). This discrepancy is likely to be because ferricinium is only a moderately large cation, so that ΔGtr(Fic+) is less positive than ΔGtr(Foc) for transfer to water. The discrepancy is not because of abnormal electrochemical behavior of the Pt/Foc, Fic+ electrode in water or other solvents. Values of E° vs nhe, H2O in a variety of solvents based on the TATB assumption are presented.  相似文献   


16.
Partial conductivities in the SrCe(Y)O3−δ system have been studied in oxidising conditions in the temperature range 923–1273 K. Compositions with variable Y content (5 and 10 at.%), Sr deficiency (3 at.%), and with the addition of Fe2O3 as sintering aid (2 mol%) were analysed. A modified Faradaic efficiency method and oxygen permeation measurements were employed to appraise the oxide-ionic transport. Oxide-ion transference numbers in air lie in the range 0.19–0.80 and decrease with increasing temperature in the range 973–1223 K. Modelling of total conductivity as a function of oxygen partial pressure (p(O2)) confirmed that protonic transport is minor under the studied conditions. SrCe0.95Y0.05O3−δ exhibits greater oxide-ion conductivity than SrCe0.9Y0.1O3−δ, indicative of dopant–vacancy association at high dopant contents. Conversely, oxygen permeability is slightly higher for SrCe0.9Y0.1O3−δ as a result of faster surface-exchange kinetics. The oxygen flux through Fe-free membranes is dominated by the bulk in low p(O2) gradients, when the permeate-side p(O2) is higher than 0.03 atm, but surface exchange plays an increasing role with increasing p(O2) gradient. Addition of Fe2O3 to SrCe(Y)O3−δ lowers the sintering temperature by 100 K but results in the formation of intergranular second phases which block oxide-ionic and electronic transport, and thus oxygen permeation. The average thermal expansion coefficients (TECs) are (10.8–11.6) × 10−6 K−1 in the temperature range 373–1373 K for all studied compositions.  相似文献   

17.
A detailed temperature variation (18–50 °C) FTIR/ATR study of sorption and desorption of water into a series of cured epoxy resins has been reported. For higher temperatures (35–50 °C) the data were modelled with a single Fickian diffusion equation, giving an increased D as the temperature increased and an activation energy (EA) in the 55–60 kJ mol−1 region. At lower temperatures (18–35 °C)—well-below the Tg—a two-stage sorption equation was needed and the apparent EA was negative. This is probably associated with changes in water clustering among the distributed ‘voids’ in the glassy polymer associated with chain relaxation at extended times. The use of D2O as a penetrant allowed diffusion coefficient measurements for highly dense epoxy matrices, where FTIR/ATR cannot detect the ν(OH) band of water over and above the residual polymer–OH groups (in the dry state). The data for the D2O studies were notably influenced by isotopic exchange; which was found to be a diffusion controlled process, even in a polymer matrix.  相似文献   

18.
Mechanical behaviour of partially stabilized zirconia crystals (PSZC) with terbia and ceria additives was investigated under bending and indentation conditions. Test specimens were oriented along the [010] direction and along the axis of crystal growth. The PSZC bending strength (σb) was dependent on the crystallographic orientation of the specimens. The specimen volume subjected to stress influenced the PSZC strength. The highest mechanical characteristics were measured for ceria-doped crystals (σb = 1.9 GPa, Klc = 11.4 MPa m1/2, Ed = 366 GPa). The failure process was studied on the Vickers indentation, with special emphasis put on the development and propagation of lateral cracks. Anisotropy of lateral cracks in the (100) plane associated with that of the elastic moduli was revealed. At the same time anisotropy of radial cracks and hardness was not found. A new version of the equation to evaluate the fracture toughness (Kcv) on the Vickers indentation was derived. The Kcv values calculated by this equation correspond to those (Klc) obtained by an SENB method.  相似文献   

19.
The glass transition of thermoplastics of different polydispersity and thermosets of different network structure has been studied by conventional differential scanning calorimetry (DSC) and temperature modulated DSC (TMDSC). The cooling rate dependence of the thermal glass transition temperature Tg measured by DSC, and the frequency dependence of the dynamic glass transition temperature T measured by TMDSC have been investigated. The relation between the cooling rate and the frequency necessary to achieve the same glass transition temperature has been quantified in terms of a logarithmic difference Δ=log10[|q|]−log10(ω), where |q| is the absolute value of the cooling rate in K s−1 and ω is the angular frequency in rad s−1 necessary to obtain Tg(q)=T(ω). The values of Δ obtained for various polymers at a modulation period of 120 s (frequency of 8.3 mHz) are between 0.14 and 0.81. These values agree reasonably well with the theoretical prediction [Hutchinson JM, Montserrat S. Thermochim Acta 2001;377:63 [6]] based on the model of Tool–Narayanaswamy–Moynihan with a distribution of relaxation times. The results are discussed and compared with those obtained by other authors in polymeric and other glass-forming systems.  相似文献   

20.
The effect of TiO2 on the grain growth of the ZnO–Bi2O3–CoO–MnO ceramic system prepared by chemical coprecipitation, was studied between 1150 and 1300 °C in air. Bi2O3 melts during firing, and then TiO2 dissolves into Bi2O3-rich liquid. TiO2 initially reacts with Bi2O3 to form Bi4Ti3O12. Above ≈1050 °C, Bi4Ti3O12 reacts with ZnO to form Zn2TiO4 spinel phase. The kinetic study of grain growth carried out using the expression GnGon=Ko·t·exp(−Q/RT) gave grain exponent (n) value as 6 and the apparent activation energy (Q) as 226.46 kJ/mol. 1.00 mol% TiO2 addition increased the grain growth exponent value from 6 to 7 and apparent activation energy with 1.00 mol% TiO2 addition was found to be 197.10 kJ/mol. The ZnO grain size gradually increases with increasing TiO2 content. Addition of TiO2 may increase the reactivity of the Bi2O3-rich liquid towards the ZnO grain, thus affecting the ZnO grain growth.  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