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.
The thermodynamic properties of the α and β polymorphs of NiMoO4 were directly investigated by calorimetry. The standard entropies of formation, Δf S ° T , of α and β were determined from measuring the molar heat capacity, C p,m, from near absolute zero (2 K) to high temperature (1380 K) by a relaxation method and differential scanning calorimetry. The standard enthalpies of formation, Δf H ° T , of α and β were determined by combining C p,m with the standard enthalpy of formation, Δf H °298, at 298 K obtained from drop solution calorimetry in molten sodium molybdate at 973 K. The standard Gibbs energies of formation, Δf G ° T , of α and β were determined from their Δ f S ° T and Δ f H ° T values. The Δ f G ° T values indicate that the polymorphic transformation from α to β occurs at 1000 K, consistent with the observed phase transformation at 1000 K.  相似文献   

2.
Beryllium nitride (Be3N2) vaporizes congruently in the range 1640° to 1960°K by the reaction Be, N2( c ) = 3Be( g ) + N2( g ). The equilibrium nitrogen partial pressure, in atmospheres, at the composition for congruent sublimation is given by the expression log P N2= [(–1.952 ± 0.038) × 104] T −1+ (6.509 ± 0.207). The measured enthalpy of decomposition (370 ± 5 kcal at 298° K) yields an enthalpy of formation for Be3N2( c ) of –136 ± 6 kcal/mole. The upper limit to the evaporation coefficient at 1600° to 2000°K can be set as 10–4 by comparison of equilibrium data to Langmuir data obtained with a sample of 18% porosity. The apparent enthalpy of activation for the reaction is 409 ± 7 kcal/mole at 1800°K for the porous Langmuir specimen. An expression is developed to predict the temperature dependence of the reduced apparent pressures in Knudsen studies of substances with low evaporation coefficients in terms of the enthalpy of activation. The variation in temperature dependence of the Langmuir measurements and Knudsen measurements with three different-sized orifices is consistent with predictions from this expression.  相似文献   

3.
Equilibrium partial pressures of SiF4 were measured for the reactions 2SiO2( c )+2BeF2( d )⇋SiF4( g )+Be2SiO4( c ) (log P siF4(mm) = [8.790 - 7620/ T ] ±0.06(500°–640°C)) and Be2SiO4( c ) +2BeF2( d )⇋SiF4( g ) +4BeO( c )(log P siF4(mm) = [9.530–9400/T] ±0.04 (700°–780°C)), wherein BeF2 was present in solution with LiF as molten Li2BeF4. The solubility of SiF4 was low (∼0.04 mol kg-1 atm-1) in the melt. The results for the first equilibrium were combined with available thermochemical data to calculate improved Δ Hf and Δ Gf values for phenacite (–497.57 ±2.2 and –470.22±2.2 kcal, respectively, at 298°K). The few measurements above 700°C for the second equilibrium are consistent with the temperature of the subsolidus decomposition of phenacite to BeO and SiO2 and with the heat of this decomposition as determined by Holm and Kleppa. Below 700°C, the pressures of SiF4 generated showed an increasing positive deviation from the expression given for the equilibrium involving Be2SiO4 and BeO. This deviation might have been caused by the formation of an unidentified phase below 700°C which replaced the BeO; it more likely resulted from a metastable equilibrium involving BeO and SiO2.  相似文献   

4.
The electrical conductivity and thermopower of Mn3O4 were measured in the temperature range 920° to 1530°C. Electrical conduction in cubic Mn3O4 is explained by the small polaron hopping of electron holes between Mn4+ and Mn3+ on octahedral sites. The concentrations of Mn4+ and Mn3+ are governed by the disproportionation equilibrium 2Mn oct 3+⇄Mn oct 4++ Mn oct 2+. This model also explains the electrical behavior of NiMn2O4 and CuMn2O4.  相似文献   

5.
The effect of curing temperature (40°, 60°, 80°C) on the hydration behavior of β-dicalcium silicate (β-C2S) was investigated. The β-C2S was obtained by decomposition of hillebrandite, Ca2(SiO3)(OH)2, at 600°C, has a specific surface area of about 7 m2/g, and is in the form of fibrous crystals. The dependence of the hydration reaction on temperature continues until the reaction is completed. The hydration is completed in 1 day at 80°C and in 14 days at 14°C. The hydration mechanism is different above and below 60°C, but at a given temperature, the reaction mechanism and the silicate anion structures of C-S-H do not change significantly from the initial to the late stages of the reaction. High curing temperature and long curing times after completion of reaction promote silicate polymerization. The Ca/Si ratio of C-S-H shows high values, being almost 2.0 above 60°C and 1.95 below 40°C.  相似文献   

6.
A technique for preparing Y1, Ba2Cu3O7-x/polymer composites showing superconducting resistive transition above liquid-nitrogen temperature has been developed. A sintered Y1Ba2Cu3O7-x disk was impregnated with a low-viscosity 2-ethylhexyl acrylate monomer containing benzoyl peroxide and dicumyl peroxide as polymerization catalysts. The impregnation of sintered disks was carried out under moderate vacuum for 30 min followed by an overnight soaking. The polymerization treatment consisted of a heat treatment at 60°C for 24 h followed by a second heat treatment at 80°C for 8 to 10 h. The superconductorlpo1ymer specimens exhibited superconducting resistive transition at ≃80 K. The scanning electron micrograph reveals excellent welting, infiltration, and polymerization of the monomer into the superconducting material.  相似文献   

7.
Viscosity and Density of Boron Trioxide   总被引:1,自引:0,他引:1  
Extensive and accurate viscosity values for molten B2O3 are reported over the continuous range from 20 poises (at 1400°C) to 1010 poises (at 318°C). The measurements were made with a wide-range rotating-cylinder viscometer. The melts were pretreated by bubbling superdry nitrogen through them to ensure a minimum water content. Above 800°C the temperature dependence of the viscosity curve obeyed the Arrhenius equation; below 800°C the curvature was smooth showing no breaks. At the low temperatures, although its curvature decreased, the viscosity curve did not obey the Arrhenius equation. Densities were determined between 411° and 1400°C. Over this range the volume expansion coefficient changed by 1 order of magnitude from 3.35 × 10−4/°C to 3.34 × 10−5/°C. The liquid volume expansion coefficient above 1200°C was smaller than that of the glass below the glass transition temperature, indicating some type of structural rearrangement in B2O3.  相似文献   

8.
The stannous-stannic equilibrium in binary alkali silicate and ternary silicate glasses was studied by equilibrating glassmelts with air at 1400°C. The Sn2+-Sn4+ equilibrium shifts more toward the oxidized state with increasing ionic radii of the alkali ions or with increasing concentration of the alkali ions in the same series of glasses. The slope of the straight lines obtained on plotting log (Sn4+)/(Sn2+)( pO2 )n/2 vs mol% R2O increased in the order Li→Na→K. In ternary silicate glasses having the base glass composition 20Na2O·10RO·70SiO2, the Sn2+-Sn4+ equilibrium shifts more toward the reduced state, with increasing bond strength between the divalent cations and the nonbridging oxygens. With increasing temperature, the equilibrium shifts more toward the reduced state.  相似文献   

9.
Free-surface and equilibrium sublimation pressure measurements indicate that the sublimation coefficient of tin selenide is essentially 1 from 726° to 975°K. Torsion-Langmuir measurements from 726° to 879°K gave the following expression for the free-surface pressure: log P (atm) = (7.021°0.161)-(1.036°0.013)104/ T. The third-law heat of sublimation calculated from these results was 51.5±0.3 kcal/mol SnSe. Similar treatment of the equilibrium pressures obtained by the torsion-effusion method from 789° to 975°K gave the following least-squares expression: log P (atm) = (7.473±0.133)-(1.069±0.012) 104/ T. The third-law ΔH208 was 52.8±0.5 kcal/mol.  相似文献   

10.
Tape-cast slurries of Ba2YCu3O7 powders offer a convenient means of preparing sintered ceramic samples for critical current density (Jc) measurements where the transport cross section is small and the current electrode areas are large. Samples were sintered from 900° to 1000°C and characterized for bulk density, grain size, phase composition, Tc, and Jc. Bulk density and grain size both increase with sintering temperature while all samples were single-phase perovskite except for those sintered at 900°C. The onset temperature for superconductivity is constant at about 93 K while the transition sharpens to R=0 at about 92 K for the densest samples. Jc rises with sintering temperature to a maximum of ∼103 A/cm2. A linear relationship between Jc and bulk density is predicted from microstructural considerations.  相似文献   

11.
Highly densified MgB2 superconductors were successfully fabricated using a spark plasma sintering (SPS) technique, and their superconductivity with respect to microstructural evolution was evaluated. Full densification with final density close to the theoretical density was achieved at a temperature of 1000°C within a total SPS processing time of 40 min. Both an MgB2 specimen sintered at 1000°C for 30 min and one sintered at 1050°C for 10 min exhibited a high critical transition temperature ( T c) similar to that of an MgB2 single crystal (39 K), and a very sharp superconducting transition width (Δ T ) less than 0.5 K. In addition, high critical current densities ( J c) of 7.7 × 105 A/cm2 in a field of 0.6 T at 5 K and of 8.3 × 104 A/cm2 in a field of 0.09 T at 35 K were obtained. These excellent superconducting characteristics of the SPS-processed MgB2 are attributed to uniformly distributed secondary MgO phase nanoparticles and well-developed dislocations in the microstructure that may act effectively as extrinsic flux pinning sites, resulting in the strong pinning force showing a high J c of 8.7 × 104 A/cm2 even in the condition of a field of 4 T at 5 K.  相似文献   

12.
High-temperature X-ray diffraction and differential thermal analyses showed that LiGa5O8 exists in two polymorphs related by the first-order transition at 1138°±3°C of the low-temperature simple-cubic form, space group (probably) O7, to the high-temperature spinel (fcc) form, space group O h 7. The transition is rapid, and the high-temperature form in pure LiGa5O8 could not be quenched to room temperature under the conditions used. However, the high-temperature polymorph can be quenched under equilibrium conditions when 40 mol% or more MgGa2O4 is present. The subsolidus equilibrium relations in the system MgGa2O4-LiGa5O8 are discussed.  相似文献   

13.
14.
The cation-defective NiAs-type polymorphs of MnCr2S4, FeCr2S4, and CoCr2S4 were prepared by high pressure-high temperature heat treatments. The magnetizations of these specimens and NiCr2S4 were studied from 1.5'K to room temperature. All phases with the NiAs-type structure were antiferromagneticalIy ordered below NBel temperatures of 70°, 200°, 300°, and 200°K, respectively. These data show a transition from ferrimagnetic to anti-ferromagnetic ordering in the transformation spinel-type ↔ NiAs-type structure in the A2+Cr2S4 compounds.  相似文献   

15.
The phase diagram of the system BaO-Fe203 was determined by X-ray diffraction, melting-point measurement, and microscopic methods. Since the reduction of Fe3+ to Fe2+ was observed by chemical analysis in the samples heated at high temperature, especially in molten samples, the samples were heated at 1 atmosphere pressure of oxygen in the temperature region in which the liquid was in equilibrium; 1 atmosphere pressure of oxygen was su5cient to restrain the reduction of Fe3+. In the temperature region of solid-solid equilibrium, the dissociation was not observed even when the samples were heated in air. BaO - 6Fe2O3 formed a solid solution with BaO.Fe203. The BaO:Fe203 ratio of the solid solution was BaO.4.5Fe203 at 1350°C. and BaO. 5.0Fe20a at 800°C. The precipitation micro-structures of each primary solid solution were observed.  相似文献   

16.
Measurements are reported for the thermal conductivity, λ(80° to 400°K), electrical resistivity, ρ(4.2° to 400°K), and absolute Seebeck coefficient, Q(6° to 400°K), of pressed and sintered uranium mononitride. The measurements between 77° and 400°K were made using an absolute longitudinal heat flow apparatus. These data and literature values for the thermal conductivity and electrical resistivity at higher temperatures were used to separate the electronic and lattice portions of the thermal conductivity. The results indicate that the lattice conductivity peaks in the range 250° to 300°K and that the high-temperature limit of the Lorenz function may be greater than the Sommerfeld value of 2.443 × 10-8 (VK-2). The electrical resistivity and the absolute Seebeck coefficient exhibit sharp slope changes near the Néel temperature, T N(∼50° to 60°K). The Seebeck coefficient has a minimum at 34°K and then rises to a local maximum at 10°K. This low-temperature peak is probably due to magnon drag. The temperature dependence of the electrical resistivity is dominated by the magnetic contribution which increases as T 2.38±0.08 between 10° and 52°K. The magnetic contribution is constant at high temperatures with an estimated value of 142 μΩ-cm.  相似文献   

17.
This paper details the investigation of the quality factor ( Q ), dielectric permittivity (ɛr) and temperature coefficient of resonant frequency (τf) of the TE01δ mode of the columbite binary niobate ceramics, with the formula MNb2O6 where M=2+ cation, in relation to their degree of sintering, microstructure and phase composition. The ceramics were made from a mixed oxide preparative route and fired over a range of temperatures from 800° to 1400°C, and most formed the columbite structure. A comprehensive study was made of the niobates containing the transition metal cations M=Mn2+, Co2+, Ni2+, Cu2+, and Zn2+, and the group II metal cations M=Mg2+, Ca2+, Sr2+, and Ba2+. All columbite niobates were found to have ɛr between 17 and 22 and negative τf values between –45 and –76 ppm/°C, and ZnNb2O6, MgNb2O6, CaNb2O6, and CoNb2O6 had high Q f values of 84 500, 79 600, 49 600, and 41 700 GHz, respectively. The Q f of MgNb2O6 was found to rise to over 95 000 GHz when heated at 1300°C for 50 h.  相似文献   

18.
The kinetics of enthalpy relaxation (recovery) at the glass transition in x K2O·(20− x )MgO·80TeO2 glasses has been examined from heat capacity measurements using differential scanning calorimetry to clarify the features of the structural relaxation in ternary TeO2-based glasses. Ternary glasses such as 10K2O·10MgO·80TeO2 show high thermal resistance against crystallization compared with binary glasses. The degree of fragility m estimated from the activation energy for viscous flow E η and the glass transition temperature T g is m = 55–62, indicating a fragile character in TeO2-based glasses. Large heat capacity changes of 43.1–48.2 J·mol−1·K−1 are also observed at the glass transition. The activation energy for enthalpy relaxation Δ H is evaluated from the cooling rate dependence of the limiting fictive temperature, and values of Δ H = 897–1268 kJ·mol−1 are obtained. Negative deviation from additivity in Δ H is also observed. Values of the Kovacs–Aklonis–Huchinson–Ramos (KAHR) parameter θ estimated from Δ H and T g are 0.33–0.42 K−1. It has been proposed that ternary glasses have more homogeneous and constrained glass structure compared with binary glasses.  相似文献   

19.
Polycrystalline bulk samples of Ti3SiC2 were fabricated by reactively hot-pressing Ti, graphite, and SiC powders at 40 MPa and 1600°C for 4 h. This compound has remarkable properties. Its compressive strength, measured at room temperature, was 600 MPa, and dropped to 260 MPa at 1300°C in air. Although the room-temperature failure was brittle, the high-temperature load-displacement curve shows significant plastic behavior. The oxidation is parabolic and at 1000° and 1400°C the parabolic rate constants were, respectively, 2 × 10−8 and 2 × 10−5 kg2-m−4.s−1. The activation energy for oxidation is thus =300 kJ/mol. The room-temperature electrical conductivity is 4.5 × 106Ω−1.m−1, roughly twice that of pure Ti. The thermal expansion coefficient in the temperature range 25° to 1000°C, the room-temperature thermal conductivity, and the heat capacity are respectively, 10 × 10−6°C−1, 43 W/(m.K), and 588 J/(kgK). With a hardness of 4 GPa and a Young's modulus of 320 GPa, it is relatively soft, but reasonably stiff. Furthermore, Ti3SiC2 does not appear to be susceptible to thermal shock; quenching from 1400°C into water does not affect the postquench bend strength. As significantly, this compound is as readily machinable as graphite. Scanning electron microscopy of polished and fractured surfaces leaves little doubt as to its layered nature.  相似文献   

20.
The role of titanium oxide in some important refractory systems was elucidated by studying selected equilibria in the system CaO-MgO-iron oxide-titanium oxide at O2 pressures of 0.21 atm (air) and 10−9 atm and under the extreme reducing conditions imposed by the presence of metallic Ti in contact with the oxide phases. Solidus relations were determined for the system CaO-MgO-TiO2 in air; 6 composition triangles were delineated, within each of which 3 crystalline phases coexist in equilibrium with liquid at a constant solidus temperature. The solidus temperatures range from 1407° to 1670°C. There is also a composition area within which MgO coexists with a Ca4Ti3O10-Ca3Ti2O7 solid solution, with solidus temperatures varying continuously from 1659° to 1670°C. Studies of reactions between MgO and titanium oxide in contact with metallic Ti in a closed system indicate that the mutual solubility between MgO and TiO at 1400°C is very small. Addition of 5 wt% TiO2 to the system CaO-MgO-iron oxide at 1500°C in air and in 10−9 atm O2 decreases the amount of iron oxide which can be absorbed by a CaO-MgO body without formation of a liquid phase; hence, titanium oxide has a strong deleterious effect on the refractoriness of such bodies.  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