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 standard entropy change ΔS° for the reduction of nonmagnetic, nonconducting oxides, MmOn(s) = mM(s) + (n2)O2(g), has been estimated as a function of m, n, and temperature T from motional entropies of oxygen molecules and vibrational entropies of solid phases. An available formula of ΔS°calc = a · m + b · n with constant a and b based on effective Debye temperatures, θM = 165 K for M and θOX = 540 K for MmOn, agrees well with the observed ΔS°obs for M2O, MO, M2O3, MO2, M2O5, and MO3 in the temperature range T = 300 – 1300 K. Possible electronic entropy corrections are applied to ΔS°calc for M2O7 and MO4.  相似文献   

2.
3.
Linear and branched bisphenol A polycarbonate (PC) samples were characterized by their average molecular weights, Mn and Mw, polydispersity degree q = Mw/Mn, and branching degree gv. The weight fraction of microgel was also determined for branched samples. The samples were amorphized and densities were measured at 23°C to obtain the values of specific volume, vsp. The dependence of vsp on molecular characteristics is described by the multivariable power function Δvsp = AspMxaqapx gvab, where Δvsp = vsp ? vsp,∞, and Asp, a, apx and ab are constants. It has been confirmed that a = ?1, apn = 0 and apw = 1. It has also been found that the branching exponent ab significantly depends on microgel content. The relationships found for PC should, in principle, be valid for other polymers. Examples based on literature data are given for linear polyethylene and polydimethylsiloxane.  相似文献   

4.
The vaporization of o-, m-, and p-dinitrobenzenes was investigated by means of the torsion-effusion method and the selected equations for vapour pressure p as a function of temperature T are:
o-dinitrobenzene: log10(patm)=(7.03±0.34)?(4270±120) KT,m-dinitrobenzene: log10(patm)=(7.66±0.28)?(4400±100) KT,p-dinitrobenzene: log10(patm)=(8.34±0.34)?(4860±120) KT
The sublimation enthalpies ΔHo(o-, 298.15 K) = (21.0 ± 0.5) kcalth mol?1, ΔHo(m-, 298.15 K) = (20.8 ± 0.2) kcalth mol?1, and ΔHo(p-, 298.15 K) = (23.0 ± 0.6) kcalth mol?1, are also derived by means of the second- and third-law treatments of the results.  相似文献   

5.
In order to elucidate the defect structure of the perovskite-type oxide solid solution La1?xSrxFeO3?δ (x = 0.0, 0.1, 0.25, 0.4, and 0.6), the nonstoichiometry, δ, was measured as a function of oxygen partial pressure, PO2, at temperatures up to 1200°C by means of the thermogravimetric method. Below 200°C and in an atmosphere of PO2 ≥ 0.13 atm, δ in La1?xSrxFeO3?δ was found to be close to 0. With decreasing log PO2, δ increased and asymptotically reached x2. The log(PO2atm) value corresponding to δ = x2 was about ?10 at 1000°C. With further decrease in log PO2, δ slightly increased. For LaFeO3?δ, the observed δ values were as small as <0.015. It was found that the relation between δ and log PO2 is interpreted on the basis of the defect equilibrium among Sr′La (or V?La for the case of LaFeO3?δ), V··O, Fe′Fe, and Fe·Fe. Calculations were made for the equilibrium constants Kox of the reaction
12O2(g) + V··o + 2FexFe = Oxo + 2Fe·Fe
and Ki for the reaction
2FexFe = FeFe + Fe·Fe·
Using these constants, the defect concentrations were calculated as functions of PO2, temperature, and composition x. The present results are discussed with respect to previously reported results of conductivity measurements.  相似文献   

6.
The mutual solubilities of {xCH3CH2CH2CH2OH+(1-x)H2O} have been determined over the temperature range 302.95 to 397.75 K at pressures up to 2450 atm. An increase in temperature and pressure results in a contraction of the immiscibility region. The results obtained for the critical solution properties are: To(U.C.S.T.) = 397.85 K and xo = 0.110 at 1 atm; (dTodp) = ?(12.0±0.5)×10?3K atm?1 at p < 400 atm and (dTodp) = ?(7.0±0.7)×10?3K atm?1 at 800 atm < p < 2500 atm; (dxodT) = ?(4.0±0.5)×10?4K?1.  相似文献   

7.
We report here a method which affords the magnitude of the characteristic parameters of a macromolecular chain starting from a polydisperse sample. A statistical treatment of viscosity measurements made on the small fractions of a GPC elution wave enables us to assign to each of them its Mn value. By numerical calculations it is then possible on the one hand to evaluate the Mark-Houwink constants, MnandMw, the f(M) function and on the other hand to estimate according to a hydrodynamic model, the unperturbed geometrical dimensions. The method is tested for polystyrene samples under conditions which allow the dissolution of many polymers, viz in tetrachloroethane (TCE) at 50° and N-methylpyrrolidone at 85°.  相似文献   

8.
The phase relationships of poly(N-vinyl-3,6-dibromo carbazole) (PVK-3, 6-Br2) were examined for four solvents, viz, o-chlorophenol, p-chloro-m-cresol, o-dichlorobenzene and bromobenzene. Upper critical solution temperatures (UCST) have been determined for solutions of PVK-3,6-Br, fractions in o-chlorophenol and p-chloro-m-cresol over the molecular weight range Mw × 10?4 = 125.0 to 4.8. The Flory temperature, θ, obtained from UCST for the PVK-3,6-Br2/o-chlorophenol and PVK-3,6-Br2/p-chloro-m-cresol systems are 66.0 and 112.9°C, respectively. The θ-temperatures were checked against molecular weight and viscosity data to determine the Mark-Houwink equations for these two theta solvents, with satisfactory agreement. The relations are
[ν] = 27.54 × 10?10 × M0.50w (o-chlorophenol, 60.0°C
[ν] = 30.24 × 10?10 × M0.50w (p-chloro-mcresol, 112.9°C
The characteristic ratio C = 〈R20/nl2 was found to be 16.6 in o-chlorophenol at 60.0°C and 17.6 in p-chloro-m-cresol at 112.9°C. The value of the characteristic ratio C of PVK-3,6-Br2 is of the same order of that for poly(N-vinyl carbazole). This indicates that the bromine atoms at the 3 and 6 (meta) positions have only an inappreciable effect on the hindering potential for rotation about the CC bond. This agreement of C for both polymers may also be taken as indicating that the effect of interaction between polar groups at the m-position on the hindering potential for rotation is small. The phase diagrams of PVK-3,6-Br2 obtained in o-dichlorobenzene and bromobenzene seem to be characteristic of organized phase structures such as those found in systems exhibiting thermoreversible gelation. Light scattering measurement on PVK-3,6-Br2 dissolved in o-dichlorobenzene, a gelation promoting solvent, and tetrahydrofuran, a very good solvent, strongly indicate that the macromolecular species in o-dichlorobenzene contain some extent supermolecular structures (aggregates, association of chain segments, etc.). These characteristic structures of PVK-3,6-Br2 in o-dichlorobenzene and bromobenzene at 25°C are also characterized by high values of the Huggins' constant k′; for tetrahydrofuran solutions, the k′ values were in the range normally found for many good solvent-polymer systems.  相似文献   

9.
The heat capacity of the solid solution Mn3.2Ga0.8N was measured between 5 to 330 K by adiabatic calorimetry. A sharp anomaly with first-order character was detected at TA = (160.5±0.5) K, corresponding to a magnetic rearrangement and a lattice expansion. No sharp anomaly was observed at Tc ≈ 260 K where the magnetic ordering takes place; instead, a smooth shoulder was detected. The thermodynamic functions at 298.15 K are Cp,mR = 15.16, SmoR = 18.57, {Hmo(T)?Hmo(0)}R = 2896 K, ?{Gmo(T)?Hmo(0)}RT = 8.85. At low temperatures the coefficient for the linear electronic contribution to the heat capacity was derived: γ = (0.031±0.003) J·K?2·mol?1. Moreover, the different contributions to the heat capacity were obtained and the electronic origin of the phase transitions was established.  相似文献   

10.
We present the heat capacities measured by adiabatic calorimetry from 6 to 350 K, and by differential scanning calorimetry from 300 to 500 K, of CsCrCl3 and RbCrCl3. A first-order transition at Tc = (171.1±0.1) K was detected for CsCrCl3. The RbCrCl3 showed at Tc = (193.3±0.1) K a transition with thermal hysteresis at temperatures just below the maximum. At T1 = (440±10) K a continuous transition was also detected. Furthermore, at TN ≈ 16 K, and for both compounds, a small bump due to magnetic long-range ordering was observed. The thermodynamic functions at 298.15 K are
  相似文献   

11.
The general equation
P = Aπi=1nvaii,
where P is a polymer property, Π is the sign of product, A is a constant, vi is the ith variable and ai is the exponent of ith variable, has been proposed for the dependence of some polymer properties on molecular weight, molecular weight distribution and long-chain branching. The data confirming the proposed equation have been taken from published theoretical and experimental papers on intrinsic viscosity, melt viscosity and glass transition temperature, as well as on viscosity of polymer solutions. Examples of application are given.  相似文献   

12.
A model theory of viscosity η for moderately concentrated polymer solutions is based on the assumption of a “local viscosity” effect and intermolecular hydrodynamic and thermodynamic interactions. It is shown that η is given by
η = ηo{1 + γc[η]}12·expHoRT1 ? aø
where γ is 0–0.4 and depends on the quality of the solvent, a varies between 0,4 and 0.8 and depends on the fraction of the “free volume” of the systems, H0 is the activation energy of the solvent and π is the polymer volume concentration. The dependence of η and “activation energy” of π and T for various molecular weights and qualities of solvents is described quantitatively. Anomalous dependences of [η] and of η on M for low polymer are obtained. An expression for η is proposed:
ηηo1 ? 2K= {1 + (1 ? 2K)c[η]}F(π)
where K is the Huggins-Martin coefficient and F(π) = 1 for most solutions when T is > Tg. For poor solvents the H vs c curve (where H is the activation energy of η of solution) has a minimum value at moderate concentrations. For good solvents, H depends slightly on the molecular weight according to an empirical equation:
H = Ho + 660α31nηηo
Expressions are given from the viscosities of solutions of miscible and also solutions of immiscible polymers.  相似文献   

13.
The structure of a KxP2W4O16 (x ? 0.4) crystal was established by X-ray analysis. The solution in the cell of symmetry P21m, with a = 6.6702(5), b = 5.3228(8), c = 8.9091(8) Å, β = 100.546(7)°, Z = 1, has led to R = 0.033 and Rw = 0.036 for 2155 reflections with σ(I)I ≤ 0.333. This structure can be described as two octahedra-wide ReO3-type slabs connected through “planes” of PO4 tetrahedra. A new structural family KxP2W2nO6n+4 can be foreseen which is closely related to the orthorhombic P4W8O32 and the monoclinic RbxP8W8nO24n+16 series.  相似文献   

14.
The thiochlorides Mo6Cl10Y (Y = S, Se, Te) have been prepared; they are isostructural with Nb6I11, space group Pccn, and have four formula units per unit cell. The X-ray structure of Mo6Cl10Se has been determined from three-dimensional single-crystal counter data and refined to a final R value of 0.053 for 3350 independent reflections. The most important result concerning this structure is a statistical distribution of the Se atom on the unit (Mo6X8) with X′ ? 78Cl + 18Se: so the compound Mo6Cl10Se must be formulated (Mo6Cl7Se)Cl42Cl22. The diamagnetic and dielectric behavior of these new thiochlorides is discussed.  相似文献   

15.
An investigation has been carried out into the effect of the fractional composition on the rheological (flow and elastic) properties of a system, using mixtures of polybutadienes with narrow molecularweight distribution (MWD). For mixtures of high-molecular-weight components, the initial Newtonian viscosity is determined by the weight-average molecular weight: η0Mαw; when low-molecular weight components are introduced, it is also determined by the MWD moment ratio. The characteristic relaxation time of a system is determined by the z-average molecular weight: θ0Mα1z, and in the general case α1α. A new model has been proposed to explain the non-Newtonian phenomenon as a consequence of the existence of a molecular-weight distribution. According to this model, as the shear rate increases the high-molecular-weight components gradually (at their critical rates) pass over to the high-elastic state. Therefore, at high shear rates, their contribution to viscous losses of a polydisperse polymer is associated with their behaviour as a viscoelastic filler in a viscous liquid.  相似文献   

16.
The kinetics of the thermal polymerization of styrene have been studied over the range 60–250°. The overall energy of activation was 86 ± 2 kJ/mole, a value identical to that obtained for the thermal polymerization of styrene in diethyl adipate. As expected, the molecular weight of the polymer decreases with increase in the temperature of the polymerization, and the ratio MwMn becomes greater than 2 for polymer formed at above 140°. The plot of log (1Mn) against (absolute temperature)?1 can be represented by two straight lines yielding 24.5 and 32,0 kJ/mole for the activation energies at temperatures below 120° and above 140°, respectively. The former value is in keeping with the molecular weight being controlled by chain transfer with monomer; the latter value would be that expected if the termination process controls the molecular weight of the polymer. Mark Houwink relationships between intrinsic viscosity and Mn and Mw have been found to apply to polymer samples when the molecular weight averages were determined by osmometry and by light scattering. However, deviations were found for low molecular weight material when measured using gel permeation chromatography. The K values were considerably lower, and the α values higher than reported in the literature.  相似文献   

17.
The cationic homopolymerization of 1-chlorobutadiene and its copolymerization with isobutene have been carried out in dichloromethane solution. The catalyst-cocatalyst pairs used were AlEt2Cl-(CH3)3CCl and AlEtCl2-(CH3)3CCl. In the latter case, the use of (CH3)3CCl was not necessary. According to experimental conditions, high molecular weight copolymers (Mn = 200,000) insoluble in CH2Cl2, or soluble low polymers (Mn = 20,000) were obtained. Homopoly-1-chlorobutadienes were soluble low molecular weight polymers (Mn = 10,000). The 1-chlorobutadiene units in the polymers are nearly exclusively 3,4 (mainly trans). Consequently, this method is not convenient to prepare elastomers with a 1,4 structure similar to that of chlorinated butyl rubber.  相似文献   

18.
The compound Th0.25 NbO3 melts congruently at 1390°C. Single crystals obtained by slow cooling from the melt are transparent and show uniaxial optical properties. A single-crystal X-ray analysis confirms the tetragonal cell found by Kovba and Trunov from a powder data and gives a = 3.90 Å and c = 7.85 Å. No systematic absence of the hkl reflections is observed on precession films. The relative intensities of the main reflections are characteristic of a perovskite-like arrangement ABO3 whose large dodecahedral A sites are only partly occupied. Several domains have been found in the perovskite-type solid solution (1 ? x) Th0.25NbO3-x NaNbO3. For 0 ? x ? 0.5 the phases have a tetragonal cell with a ? a0 and c ? 2a0 as in pure Th0.25 NbO3. When 0.6 ? x ? 0.8 the corresponding phases crystallize with a small cubic cell (a0 ? 3.9Å), while phases with 0.9 ? x ? 1 have an orthorhombic cell (a ? 212a0, b ? 212a0, c ? a0).  相似文献   

19.
In this paper, we present dielectric results for samples of POE (polyethyleneoxyde) divided into three classes. Compounds of low molecular weight (Mw < 1000) show very weak absorption at lower temperatures. At higher temperatures (below the melting point), there appears a very important peak (stronger for lower Mw) which corresponds to the supercooled phase. Compounds of intermediate molecular weight (1000 ≤ Mw ≤ 105) show a β relaxation near 250 K (bisMw ? 3000) which is due to the movement of chain-ends in the amorphous phase. This process increases with the importance of that part of the structure formed of shorter lamellae, thus explaining the marked diminution of this absorption for higher molecular weight compounds (the shorter lamellae cannot be obtained for Mw > 4000). The α relaxation (obtained for T > 300 K) is perturbed by an important conduction; however, we have found a peak and a shoulder at lower frequencies, due to shorter and longer lamellae respectively. The shoulder is also present for high molecular weight compounds (Mw ≥ 0.1–106). At lower temperatures, they display a very broad γ absorption, superposed at higher frequencies with a β relaxation. The nature of this β relaxation is different from that observed for the intermediate molecular weight compounds.  相似文献   

20.
It is well known that the apparent specific volume η2 of a polymer may be expressed by the following relationship: η2= ηm + K/Mn where M?n is the number-average molecular weight of the polymer, ηm the specific volume of the infinite polymer, and K a constant. We have shown that this relationship is valid for low molecular weight polystyrenes (Mn < 4·104) with different end-groups, independently of the nature of the solvent. The K values (and variations with the solvent and with the nature of the end-groups) may be predicted through simple calculations proposed here. We conclude that ηm does not represent the specific volume of the infinite polymer, since we observe a rapid decrease of η2 for increasing M (when Mn < 4·104). The decrease is much greater than expected from the relationship η2 = ? (1/M).  相似文献   

Cp,mRSmoR{Hmo(T)?Hmo(0)}RK?{Gmo(T)?Hmo(0}RT
CsCrCl315.3826.493503.214.735
RbCrCl315.7625.993556.814.384
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