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.
Pure aluminium containing about 200?at.ppm Fe in solution is shown to creep about 106 times slower at 200°C than the same aluminium containing a negligible amount of iron in solution. The high creep resistance of the Al–200?at.ppm?Fe alloy is attributed to the presence of subgrain boundaries containing iron solute atoms. It is proposed that the opposing stress fields from subgrain boundaries and from the piled-up dislocations during creep are cyclically relaxed, by iron solute diffusion, to allow climb of the lead dislocation in the pile-up. The mechanism is a form of mechanical ratcheting. The model is applied to Al–Fe alloys and correctly predicts that the creep rate is controlled by the rate of iron solute diffusion and by a temperature dependence equal to the activation energy for iron diffusion, namely Q c?=?221?kJ?mol?1. Basic creep studies on solid-solution alloying with solute atoms that diffuse slowly in the lattice of aluminium (e.g. manganese, chromium, titanium and vanadium) appear worthy of study as a way of enhancing creep strength and of understanding creep mechanisms involving solute-atom-containing subgrain boundaries.  相似文献   

2.
B. Amami  M. Addou  F. Millot  A. Sabioni  C. Monty 《Ionics》1999,5(5-6):358-370
Measurements of18O self-diffusion in hematite (Fe2O3) natural single crystals have been carried out as a function of temperature at constant partial pressure aO 2=6.5·10?2 in the temperature range 890 to 1227 °C. The aO 2 dependence of the oxygen self-diffusion coefficient at fixed temperature T=1150 °C has also been deduced in the aO 2 range 4.5·10?4 - 6.5·10?1. The concentration profiles were established by secondary-ion mass spectrometry; several profiles exhibit curvatures or long tails; volume diffusion coefficients were computed from the first part of the profiles using a solution taking into account the evaporation and the exchange at the surface. The results are well described by $$D_O \left( {{{cm^2 } \mathord{\left/ {\vphantom {{cm^2 } s}} \right. \kern-\nulldelimiterspace} s}} \right) = 2.7 \cdot 10^8 a_{O_2 }^{ - 0.26} \exp \left( { - \frac{{542\left( {{{kJ} \mathord{\left/ {\vphantom {{kJ} {mol}}} \right. \kern-\nulldelimiterspace} {mol}}} \right)}}{{RT}}} \right)$$ From fitting a grain boundary diffusion solution to the profile tails, the oxygen self-diffusion coefficient in sub-boundaries has been deduced. They are well described by $$D''_O \left( {{{cm^2 } \mathord{\left/ {\vphantom {{cm^2 } s}} \right. \kern-\nulldelimiterspace} s}} \right) = 3.2 \cdot 10^{25} a_{O_2 }^{ - 0.4} \exp \left( { - \frac{{911\left( {{{kJ} \mathord{\left/ {\vphantom {{kJ} {mol}}} \right. \kern-\nulldelimiterspace} {mol}}} \right)}}{{RT}}} \right)$$ Experiments performed introducing simultaneously18O and57Fe provided comparative values of the self-diffusion coefficients in volume: iron is slower than oxygen in this system showing that the concentrations of atomic point defects in the iron sublattice are lower than the concentrations of atomic point defects in the oxygen sublattice. The iron self-diffusion values obtained at T>940 °C can be described by $$D_{Fe} \left( {{{cm^2 } \mathord{\left/ {\vphantom {{cm^2 } s}} \right. \kern-\nulldelimiterspace} s}} \right) = 9.2 \cdot 10^{10} a_{O_2 }^{ - 0.56} \exp \left( { - \frac{{578\left( {{{kJ} \mathord{\left/ {\vphantom {{kJ} {mol}}} \right. \kern-\nulldelimiterspace} {mol}}} \right)}}{{RT}}} \right)$$ The exponent - 1/4 observed for the oxygen activity dependence of the oxygen self-diffusion in the bulk has been interpreted considering that singly charged oxygen vacancies V O ? are involved in the oxygen diffusion mechanism. Oxygen activity dependence of iron self-diffusion is not known accurately but the best agreement with the point defect population model is obtained considering that iron self-diffusion occurs both via neutral interstitals Fe x i and charged ones.  相似文献   

3.
The diffusion coefficient of nickel in cold-worked carbon steel was determined with the diffusion couple method in the temperature range between 320 and 450 °C. Diffusion couple was prepared by electro-less nickel plating on the surface of a 20% cold-worked carbon steel. The growth in width of the interdiffusion zone was proportional to the square root of diffusion time to 12,000 h. The diffusion coefficient (DNi) of nickel in cold-worked carbon steel was determined by extrapolating the concentration-dependent interdiffusion coefficient to 0% of nickel. The temperature dependence of DNi is expressed by DNi = (4.5 + 5.7/?2.5) × 10?11 exp (?146 ± 4 kJ mol?1/RT) m2s?1. The value of DNi at 320 °C is four orders of magnitude higher than the lattice diffusion coefficient of nickel in iron. The activation energy 146 kJ mol?1 is 54% of the activation energy 270.4 kJ mol?1 for lattice diffusion of nickel in the ferromagnetic state iron.  相似文献   

4.
The γ-TiAl intermetallic compound with suitable alloying additions has shown considerable promise as a material for high-temperature applications. Diffusion studies in this alloy system are useful in assessment of their creep behaviour and structural stability in service conditions. Tracer diffusion coefficients of 51Cr and 54Mn in a γ-TiAl intermetallic compound containing 54.1 at. % aluminium were determined in the temperature range from 1095 to 1470?K. The temperature dependence of both the diffusing species follows a linear Arrhenius behaviour and can be expressed as D Cr?=?4.4?×?10?3exp(?350?kJ?mol?1/RT)?m2?s?1 and D Mn?=?1.2?×?10?3?×exp(?326?kJ?mol?1/RT)?m2?s?1. The data are analysed on the basis of empirical correlations between the diffusion and melting parameters applicable for conventional mono-vacancy diffusion mechanism in metals. It is concluded that impurity diffusion in γ-TiAl occurs through the migration of thermal vacancies via nearest-neighbour or next-nearest neighbour jumps.  相似文献   

5.
G3(MP2)//B3LYP calculations have been carried out on trans‐ and cis‐decalin, and their mono‐, di‐, tri‐, and tetraoxa‐analogs. The main purpose of the work was to obtain enthalpies of formation for these compounds, and to study the relative stabilities of the cistrans and positional isomers of the various (poly)oxadecalins. Comparison of the computational enthalpies of formation with the respective experimental ones, known only for the decalins and 1,3,5,7‐tetraoxadecalins, shows that in both cases the computational values are more negative than the experimental ones, the deviations being ?5 to ?7 kJ mol?1 for the decalins and ?12 to ?17 kJ mol?1 for the 1,3,5,7‐tetraoxadecalins. The respective computational enthalpies of cistrans isomerization, however, are in excellent to satisfactory agreement with the experimental data. The cistrans enthalpy differences vary from +11.0 kJ mol?1 for decalin to ?15.4 kJ mol?1 for 1,4,5,8‐tetraoxadecalin. Low relative enthalpy values were also calculated for the cis isomers of 1,8‐dioxadecalin (?3.7 kJ mol?1), 1,3,6‐trioxadecalin (?4.6 kJ mol?1), 1,3,8‐trioxadecalin (?9.7 kJ mol?1), 1,4,5‐ trioxadecalin (?5.6 kJ mol?1), 1,3,5,8‐tetraoxadecalin (?7.3 kJ mol?1), and 1,3,6,8‐tetraoxadecalin (?14.5 kJ mol?1). Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

6.
Abstract

Kinetic and equilibrium data for 1-EtIm binding to cyt c at temperature range of 303–319K have been determined at pH 7.0 by using 1HNMR method. Thermodynamic values (ΔH°=39.5 kJ mol?1, ΔS° = 154 J mol?1K?1 and Ea = 142 kJ mol?1) are obtained from Van't Hoff and Arrhenius's relations. Some hyperfine shifted resonances of l-EtIm cyt c have been assigned for the first time using 1D saturation transfer experiments. The origin of the asymmetric spin density distribution in heme groups of 1-EtIm cyt c and the reason of low affinity of cytochrome c for 1-EtIm are also discussed move toward upfield from original position. However, Te-2, 4 substituents in cyt c are more electron withdrawing than the propionic acid side chains, which lead to the 5-methyl group going toward downfield.  相似文献   

7.
Using a novel diffusion-evaporation method, the self-diffusion coefficient of manganese in manganous sulphide has been determined as a function of temperature (1073–1373 K) in equilibrium with the metallic phase. It has been shown that the activation energy of this process at constant sulphur activity amounts to 269 kJ/mol and the self-diffusion coefficient is the following function of temperature and sulphur vapour pressure: DMn=0.252 P?16S2 exp(-269kJ/mol/RT). Diffusion of Mn2+ cations in Mn1+yS proceeds via the interstitialcy mechanism and the activation enthalpy of successive jumps of these defects, δHm is equal to 118 kJ/mol. It has been demonstrated that the mobility of interstitial cations in the Mn1+yS lattice does not depend on their concentration and the diffusion coefficient of these defects has the following function of temperature: Di=0.759 exp(-118 kJ/mol/RT).  相似文献   

8.
Antiferromagnetic Fe0.9S forms a ferrimagnetic superlattice in a narrow temperature range centred at 475 K (γ-transition). We have made a study in single crystals by specific heat, Mössbauer spectroscopy and magnetic susceptibility of the magnetic structure and kinetics and thermodynamics of formation of the ferrimagnetic component.The antiferromagnetic alignment and the spontaneous moment lie in the c-plane. The magnetic anisotropy energy is of the order of 0.4 K per Fe ion. Above 475 K and below TN ≈ 598 K there is a change in ordering of the vacancies and a new antiferromagnetic structure is formed. At lower temperatures our value (100 kJ mol?1) of the activation energy of diffusion of iron vacancies, determined from magnetic susceptibility, is in quite good agreement with the value (88 kJ mol?1) obtained from radiotracer measurements by Condit et al.  相似文献   

9.
Abstract

The hydrogen bonded complexes between N-methylsuccinimide and phenols (pKa = 10.2 → 6) are investigated by infrared spectrometry. The thermodynamic parameters for the 1–1 complexes are determined in carbon tetrachloride-. The formation constants at 298 K range from 15 to 150 dm3 mol?1, the enthalpies of complex formation from - 20 to - 30 kJ mol?1, the changes of entropy from - 22 to - 40 J K?1 mol?1 and the frequency shifts of the v(OH) stretching vibration from 170 to 340 cm?1. The complexes are weaker than those involving the monocarbonyl bases. The decrease of the force constant of the bonded carbonyl group ranges from 0.48 to 0.65 N cm?1. The force constant of the free C=O group slightly increases upon complex formation, in agreement with the cooperatively theory.  相似文献   

10.
The isomeric 1- and 2-naphthyl acetates (acetoxynaphthalenes) are at the confluence of diverse concepts, techniques and classes of organic compounds. Summing the results of literature measurements of the enthalpy of formation of their solids and of our new sublimation enthalpies reported herein, we derive gas phase enthalpies of formation of ?209.9 ± 1.4 and ?213.3 ± 1.3 kJ mol?1 respectively. This corresponds to 2-naphthyl acetate being more stable than its 1-isomer by 3.4 ± 1.9 kJ mol?1. We also performed MP2(full)/6-31G(d) calculations on these species, resulting in enthalpies of formation of ?212.9 ± 3.9 and ?212.2 ± 3.9 kJ mol?1 for 1- and 2-naphthyl acetate and a difference of ?0.7 kJ mol?1 respectively in satisfactory agreement with the above experimental results.  相似文献   

11.
Heat capacities of [Fe(phen)2(NCS)2] and [Fe(phen)2(NCSe)2] were measured between 135 and 375 K. A heat capacity anomaly due to the spin-transition from low-spin 1A1 to high-spin π2 electronic ground state was found at 176·29 K for the SCN-compound and at 231·26 K for the SeCN-compound, respectively. Enthalpy and entropy of transition were determined to be ΔH = 8·60 ± 0·14 kJ mol?1 and ΔS = 48·78 ± 0·71 J K?1 mol?1 for the SCN-compound and ΔH = 11·60 ± 0·44 kJ mol?1 and ΔS = 51·22 ± 2·33 J K?1 mol?1 for the SeCN-compound. To account for much larger value of ΔS compared with the magnetic contribution, we suggest that there is significant coupling between electronic state and phonon system. We also present a phenomenological theory based on heterophase fluctuation. Gross aspects of magnetic, spectroscopic, and thermal behaviors were satisfactorily accounted for by this model. To examine closely the transition process, infrared spectra were recorded as a function of temperature in the range 4000 ? 30 cm?1. The spectra revealed clearly the coexistence of the 1A1, and the 5T2 ground states around Tc.  相似文献   

12.
The temperature dependent field ionization mass spectrometry method combined with ab initio calculations was used to determine the interaction energies and the structures of 9-methylguanine-acrylamide dimers. Acrylamide mimics the side chain amide group of the natural amino acids asparagine and glutamine. The experimental enthalpy of the dimer formation derived from the van't Hoff plot is ?59.5 ± 3.8 kJ mol?1. The value is higher than interaction energies between acrylamide and other nucleic acid bases which were determined to be ?57.0 for 1-methylcytosine, ?52.0 for 9-methyladenine, and ?40.6 kJ mol?1 for 1-methyl-uracil. In total, eight hydrogen bonded dimers formed by the three lowest energy 9-methylguanine tautomers and acrylamide were found in the quantum chemical calculations performed at the DFT/B3LYP/6-31++G?? and MP2/6-31++G?? levels of theory. The relative stability and the interaction energies of the dimers were calculated accounting for the basis set superposition error and the zero-point vibrational energy correction. The lowest energy dimer found in the calculations is formed by acrylamide (Ac) with the keto tautomer of 9-methylguanine (Gk). It is stabilized by two intermolecular H bonds, C6=O(Gk) · · · H—N(Ac) and Nl—H(Gk) · · ·O(Ac), and it is more stable than the second lowest energy dimer by ≈ 25 kJ mol?1. The calculated interaction energies of the lowest energy 9-methylguanine-acrylamide dimer are ?65.0 kJ mol?1 and ?67.7 kJ mol?1 at the MP2 and DFT levels of theory, respectively. The experimental enthalpy of the dimer formation is in good agreement with both the calculated interaction energies of the GkAc dimer and much higher than the interaction energies calculated for all other 9-methylguanine-acrylamide dimers. This proved that only one dimer was present in the experimental samples. To verify whether acrylamide is a good model of the amino acid-amide group, we performed direct calculations of the 9-methylguanine-glutamine dimers at the same levels of theory as used for the complexes involving acrylamide. The interaction energies found for the lowest energy 9-methylguanine-glutamine dimer are ?65.1 kJ mon?1 (MP2/6-31++G??) and ?66.2 kJ mol?1 (DFT/B3LYP/6-31++G??) and these values are very close (within 0.5 kJ mol?1) to the interaction energies obtained for the 9-methylguanine-acrylamide dimers.  相似文献   

13.
The rotational and translational dynamics of benzene adsorbed in Na-mordenite have been studied by incoherent quasi-elastic neutron scattering. The measurements were performed at two benzene coverages at 300, 400 and 450 K. The observed quasi-elastic broadenings are described by a uniaxial rotational model about the six-fold axis of benzene. The mean time between successive jumps, at 300 K, is τ=1.45 × 10?12 s at low coverage and 2.05×10?12 s at high coverage. The correlation times follow an Arrhenius law with EA=4.51 kJ mol?1, at both coverages. The translational diffusion coefficient has been measured at 300 K and was found to be 0.67 × 10?6 cm2s?1.  相似文献   

14.
α-spectrometry was used in order to measure the diffusion of U in bulk α-Ti in the temperature range 863–1123?K (540–850?°C). A straight Arrhenius plot was found, giving diffusion parameters Q?=?297?kJ/mol and D 0?=?5?×?10?3?m2/s, which are similar to the α-Ti self-diffusion ones, when measured in Ti samples with a similar impurity content than presently. This behaviour is compatible with the hypothesis of U diffusing via a vacancy-assisted mechanism in the α-Ti lattice and contrasts with older results in which the activation energy is almost a third the self-diffusion one, even lower than the vacancy formation energy.  相似文献   

15.
Ball milled nanocrystalline AlMg4.8 powder was investigated in terms of hardening and thermal stability. The validity of the Hall–Petch relation was confirmed down to the minimum grain size of ~44 nm. Prolonged milling in the range of the minimum grain size still increased the hardness. This development is discussed in terms of contamination effects and the influence of full and partial dislocations. Concerning thermal stability, recovery processes occur in the range of 100–230°C, whereas substantial grain growth starts at a temperature of ~250°C. The enthalpy release for recovery was detected to be ~39 J mol?1 and ~208 J mol?1 for grain growth. Dynamic strain ageing was indicated by an activation energy for recovery of Q?~?120 kJ mol?1. The activation energy of grain growth was calculated by means of the Kissinger theory (Q?=?200–210 kJ mol?1) and using the results of static grain growth (Q?=?204 kJ mol?1).  相似文献   

16.
Abstract

The water exchange reaction of [Pd(dien)H2O]2+ (dien = diethylenetriamine) was studied as function of temperature (268-308 K) and pressure 0.1-197 MPa) using 17O NMR techniques. The rate and activation parameters are: kcx = 5100 s?1 at 298 K; ΔH# =38 kJ mol?1; ΔS# = -47 JK?1 mol?1; ΔV# = -2.8 cm3 mol?1 at 296 K. The results are discussed in reference to solvent exchange data for other Pt(II) and Pd(II) complexes, and are interpreted in terms of an associatively activated substitution process.  相似文献   

17.
Dongmei Wu 《Ionics》2012,18(6):559-564
Li4Ti5O12 anode was successfully synthesized by solid-state method. X-ray diffraction and scanning electron micrographs show that Li4Ti5O12 prepared by solid-state method has a purity phase with a uniform particle size in the range of 0.5?C1???m. Cyclic voltammogram reveals that there is a big irreversible capacity for the first cycle. Li4Ti5O12 shows a stable cycling stability at 1?C rate. After 152 cycles, the discharge capacity is 213?mAh?g?1, which keeps 93% of it at the second cycle. Electrochemical impedance spectroscopy shows that the resistance of charge-transfer of Li4Ti5O12 electrode decreased with increasing the storage temperatures, and the lithium diffusion coefficient is increased with increasing the storage temperatures, revealing that the kinetics of Li+ and electron transfer into the electrodes were much faster at high temperature than that at low temperature. The apparent activation energy of the charge transfer and lithium diffusion can be calculated to be 33.1 and 27.3?kJ?mol?1, respectively.  相似文献   

18.
The effect of alloy surface roughness, achieved by different degrees of surface polishing, on the development of protective alumina layer on Fe-10 at.% Al alloys containing 0, 5, and 10 at.% Cr was investigated during oxidation at 1000 °C in 0.1 MPa oxygen. For alloys that are not strong Al2O3 formers (Fe-10Al and Fe-5Cr-10Al), the rougher surfaces increased Fe incorporation into the overall surface layer. On the Fe-10Al, more iron oxides were formed in a uniform layer of mixed aluminum- and iron-oxides since the layer was thicker. On the Fe-5Cr-10Al, more iron-rich nodules developed on an otherwise thin Al2O3 surface layer. These nodules nucleated preferentially along surface scratch marks but not on alloy grain boundaries. For the strong Al2O3-forming Fe-10Cr-10Al alloy, protective alumina surface layers were observed regardless of the surface roughness. These results indicate that the formation of a protective Al2O3 layer on Fe-Cr-Al surfaces is not dictated by Al diffusion to the surface. More cold-worked surfaces caused an enhanced Fe diffusion, hence produced more Fe-rich oxides during the early stage of oxidation.  相似文献   

19.
Proton nuclear magnetic resonance relaxation times and linewidth measurements have been made on five polycrystalline organic compounds, triethylenediamine, 3-azabicyclononane, norbornane, norbornylene and norbornadiene. Measurements for each sample were made throughout the plastic crystal phase. The results are analysed in terms of molecular motion. Correlation times τ and activation enthalpies for translational self-diffusion of molecules are evaluated: triethylenediamine τ=7·6×10?19 exp (96·4/RT)s, 3-azabicyclononane τ=1·7×10?16 exp (83·6/RT)s, norbornane for 131K<T<306 K, τ=4·6×10?15 exp (54·5/RT)s for 306K<T<360K, τ=1·1×10?16 exp (64·8/RT)s, norbornylene, τ=4·×10?15 exp (48·6/RT)s and norbornadiene τ=6·8×10?15 exp (39·9/RT)s, where R is the gas constant in units of kJ K?1mol?1. The results and mechanism of diffusion are discussed in relation to the thermodynamic properties of the materials.  相似文献   

20.
High pressure/temperature annealing experiments are used to determine diffusivities of H+ and D+ in non-stoichiometric spinel, a low-pressure analogue for nominally anhydrous minerals in Earth’s mantle. Data are fitted to the following Arrhenius law: Diffusivity (m2/s)?=?4?±?1?×?10?12 exp(?54?±?2 kJ?mol?1/RT). At low temperatures, H+ and D+ diffusion in non-stoichiometric spinel is charge balanced by flux of O vacancies, with infrared data consistent with protonation of both octahedral and tetrahedral O–O edges in non-stoichiometric spinel, and additional fine structure due to Mg–Al mixing and/or coupling of structurally incorporated H+ with cation vacancies. Absence of changes in the fine structure of O–H absorption bands indicates that H+ can become locally coupled and uncoupled to other defects during bulk diffusion. As such, proton conductivity in spinel group minerals, arising from faster flux of uncoupled H+, can only be calculated from H+ mobility data if the extent of defect coupling is constrained.  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