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.
Lithium salts of two polyanionic addition polymers containing alkyl sulphonic acid and perfluoroalkyl carboxylic acid side groups were prepared. Blends of these polymers were formed with poly(ethylene oxide) (PEO). The blend containing alkyl sulphonate units showed some phase separation but this was not observed for the blend containing perfluoroalkyl carboxylate groups. In the latter case a comparatively high conductivity of ~10?5 Ω?1 cm?1 at 374 K was obtained. The anionic units in these blends are expected to be virtually immobile. Complexes formed from PEO and the Li-salt of hexafluoroglutaric acid had similar high ionic conductivities and there are grounds for supposing that the anions in these complexes may also be substantially immobilized. In addition, conductivity values were obtained for some PEO complexes containing lithium salts of some monobasic acids and it was found that the complex formed from the Li-salt of the strongest acid gave the highest conductivity (~4 × 10?4 Ω?1 cm?1 at 373 K for a PEO-LiSO3CF3 complex).  相似文献   

2.
F. Viras  T.A. King 《Polymer》1984,25(7):899-905
Low frequency Raman depolarized spectra from poly(methyl methacrylate) have been determined at temperatures between 293 K and 85 K in the frequency region 5 cm?1<v?<2000 cm?1. All the reduced intensity spectra resolve into two bands: a broad and intense band at 91 cm?1 and a narrower and weaker band at 22 cm?1. The Raman intensities do not depend on temperature which implies first order Raman scattering. Room temperature low frequency Raman spectra were also recorded from poly(ethyl methacrylate) and poly(-n-butyl methacrylate). In the two PMMA substitutes the high frequency band is shifted to lower values as the side group mass is increased and show a Debye-like character. The lower frequency band cannot be resolved in the spectrum of PEMA and PnBMA. The two bands at 91 cm?1 and 22 cm?1 are attributed to side group and backbone motions respectively. By studying PMMA samples of different tacticity no dependence of the band frequency on chain configuration has been detected. The density of states g(Ω) obtained through the low frequency Raman spectrum has been used to calculate the specific heat Cv. The values of Cv obtained are lower than those measured calorimetrically but varied with temperature in a similar manner in the region 0.8<T<4K.  相似文献   

3.
M. Kajiwara  M. Hashimoto  H. Saito 《Polymer》1973,14(10):488-490
Chelating polymers containing copper, nickel and cobalt have been formed from cyclophosphazene thiocarbamate trimer and copper, nickel or cobalt ions. These polymers obtained from the reaction are amorphous and the values of electron conductivity are 2·7 × 1011Ω-cm, 3·6 × 1013Ω-cm and 1·5 × 1013Ω-cm for the Cu, Ni and Co polymers respectively. The Cu polymer is the most thermally stable on heating to 500°C in air.  相似文献   

4.
F. Viras  T.A. King 《Polymer》1984,25(10):1411-1414
Low frequency excitations in amorphous polycarbonate-bisphenol A have been studied by Iaser Raman spectroscopy in the frequency region 5 cm?1 < Δv? < 200 cm?1. Depolarized spectra were recorded at temperatures between ambient and 85 K, the reduced intensity spectra show no temperature dependence. Two bands are resolved in the reduced intensity spectrum: a strong band around 96 cm?1 and a weak band around 45 cm?1; these are attributed to the effects of longitudinal and transverse phonon waves originating in backbone motion. The low frequency Raman intensities provide information on the density of state g(ω) from which the specific heat, Cv, of polycarbonate has been calculated. This is found to vary with temperature in a manner similar to the calorimetrically measured Cv in the low temperature range ~1 K < T < 4 K.  相似文献   

5.
This article reports the synthesis and free radical polymerization of ortho-vinylbenzophenone. The glass transition temperature Tg of the homopolymer is 136°C. The products synthesized appeared to be atactic and amorphous. The Mark-Houwink constants for poly (o-vinylbenzophenone) in tetrahydrofuran are K = 4.2 × 10?2 cm3 g?1 and a = 0.765. The pre-exponential constant under theta conditions, Kθ, is estimated to be 5.93 × 10?2 cm3 g?1. The ratio of unperturbed dimensions of the actual polymer and free rotating analogue chain is 3.93, which is almost double that of polystyrene. The Flory-Huggins interaction parameter for poly (o-vinylbenzophenone)tetrahydrofuran is 0.48 at room temperature. The kpk12t ratio at 60°C is 1.1 × 10?2l12mol?12s?12. In free radical copolymerizations with styrene at 70°C, r1 (o-vinylbenzophenone) = 1.216, r2 = 0.751. This copolymerizations is virtually random.  相似文献   

6.
The crosslinking of poly(vinyl alcohol) beads was performed by γ-ray irradiation. The maximum molecular size, Mlim, which can permeate through beads, was obtained by eluting poly(ethylene glycol) through a column packed with beads. The root-mean-square of the end-to-end distance (r2)12 and that of the hydrodynamic radius (s2)12 were also calculated. Meanwhile, the average size of the network of the beads, rc, was estimated from the average molecular weight between crosslinkages, Mc, which was derived from the glass transition temperatures of crosslinked and non-crosslinked poly(vinyl alcohol) beads. rc values of poly(vinyl alcohol) beads decreased as the total dose of γ-irradiation increased when dry poly(vinyl alcohol) beads were γ-irradiated, while the values were almost constant when poly(vinyl alcohol) beads swollen in water were γ-irradiated. Further, linear relations were obtained among rc, (r2)12, and (s2)12 covering the wide range of total dose.  相似文献   

7.
Incoherent neutron scattering spectra for polypropylene and its partially deuterated analogue allow identification of the methyl torsion at 230 cm?1. Comparison of the experimental spectra with the calculated densities of states shows discrepancies with the expected intensity of the methyl torsional band. Results from a stretch-oriented sample arranged so that the wave vector Q is first parallel (Q) and then perpendicular (Q) to the helical chain axis indicate that the torsion is more intense and probably has a small frequency dispersion around 240 cm?1 for Q and is weaker with a broad dispersion centred at 220 cm?1 for Q.  相似文献   

8.
Dense tungsten hemicarbide specimens were used to study the bulk and grain boundary diffusion of 14C into W2C at temperatures of 1200 to 2000°C. The bulk diffusion coefficient is given by:
Dv =18.3exp?91,500RTcm2s?1
The grain boundary diffusion coefficient is represented by the expression:
PGB =1.8 10?4exp?68,880RTcm2s?1
A comparison is give with preceding results on other carbides.  相似文献   

9.
B.T. Kelly 《Carbon》1974,12(5):535-541
A calculation is presented of the elastic constant C33 of a graphite crystal as a function of temperature up to 2500 K, taking into account the anharmonic contribution and the changes in interlayer interactions due to the large lattice thermal expansion. Parametric variations in the theory show that the anharmonic contribution to C33 depends principally on the parameter (?2C33?e2zz) Comparison of theoretical results with the experimental data, which is mainly from neutron scattering experiments, shows that the data can be accounted for if (?2C33?e2zz) lies in the range 7–10 × 1013 dynes/cm2. A theoretical estimate of (?2C33?e2zz) based on Lennard-Jones potentials between atoms in adjacent basal planes gives a value of 9·07 × 1013 dynes/cm2.  相似文献   

10.
D.R. Dugwell  P.J. Foster 《Carbon》1973,11(5):455-467
The rates of deposition of carbon on alumina surfaces and on soot particles, have been measured in a pilot scale tubular reactor in which cold methane was mixed with combustion products at 1920°K. A hard grey metallic film of carbon, quite free of soot, was deposited on alumina surfaces for initial methane concentrations between 12 and 24 per cent. An induction period of slow growth rate, before a film covered the surface completely, was followed by a constant growth rate. Measured growth rates were from 0·06 × 10?6 to 1·43 × 10?6 g/cm2 sec of carbon on alumina at 1270°K to 1450°K, and from 0·1 × 10?4 to 1·14 × 10?4 g/cm2 sec on soot particles at 1370°K to 1700°K. Methane decomposition rates were much higher than predicted by the unimolecular mechanism indicating a predominance of radical reactions. Carbon deposition rates were related to the mole fraction, χ, of hydrocarbons in the gas which bear more than three carbon atoms per molecule, by, m?f = 1·0 × 102 n.χ. exp (?42,300/RTf), g/cm2sec for carbon film, m?s = 4·6 × 103 exp (? 46,100/RTg), g/cm2 sec for soot. A precoat of soot increased the growth rate of film carbon by 1·8 to 7·8 times yielding a hard adherent dull brown film  相似文献   

11.
The dimensions of both cyclic and linear poly(dimethyl siloxanes) in dilute solution in benzene-d6 have been measured by small-angle neutron scattering. The mean-square radii of gyration of the linear polymers are consistent with values predicted from published data, including experimental molar cyclization equilibrium constants. The average dimensions of the cyclic poly(dimethyl siloxanes) in fractions containing z-average numbers of bonds n?z in the range 130 < n?z < 550, were found to be considerably smaller than those of the corresponding linear polymers. The neutron scattering results give a value for the ratio of the z-average radii of gyration for linear and ring poly(dimethyl siloxanes) (containing the same number of monomer units) 〈s2z,l<s2z,r = 1.9 ± 0.2. This ratio may be compared with the value of 2.0 predicted theoretically for ‘flexible’ high molecular weight linear and cyclic polymers, unperturbed by excluded volume effects.  相似文献   

12.
A sample of poly (trans-1,4-cyclohexylene-dimethylene-oxymethylene oxide) (PTCDM) was synthesized by condensation of trans-1,4-cyclohexane dimethanol and paraformaldehyde using p-toluene sulphonic acid as catalyst. A fraction having Mn=6500 and a melting point of 86°C was isolated and purified; its n.m.r. spectrum does not change with temperature in the range 20°–50°C which indicates a rigid distribution of methylene substituents in the cyclohexane ring; its dipole moment, measured in benzene solution at several temperatures between 20° and 60°C, yielded values of Dn=2]nm2=0.17–0.21 and a temperature coefficient dln {gm2}dT = 5.5 × 10?3K?1, similar to those reported in the literature for acyclic polyformals. Agreement between experimental and calculated (using rotational isomeric states theory) values is satisfactory.  相似文献   

13.
For solutions of polystyrene (M=105–106 g mol?1), intrinsic viscosities [η] have been measured at 34.5°C, which is the θ temperature for the polymer in cyclohexane. The solvents comprised cyclohexane in admixture with a thermodynamically good solvent, 1,2,3,4-tetrahydronaphthalene (tetralin, TET) over the whole range of solvent composition. From an assessment of several extrapolation procedures, a value of 85 × 10?3(±1 × 10?3) cm3g?32mol12 was obtained for Kθ (in the relationship [η] = KθM12α3, where α is the expansion factor), thus yielding 0.681 A? g?12mol12, 2.25 and 10.2 for the unperturbed dimensions, steric factor σ and characteristic ratio C respectively. The value of Kθ was independent of solvent composition despite the finite excess free energy of mixing for the solvent components alone, which has been asserted elsewhere to affect Kθ. The present results, in conjunction with previous ones relating to 98.4°C, indicate a value of ?0.89 × 10?3 deg?1 for the temperature coefficient of the unperturbed dimensions.  相似文献   

14.
Thomas C. Amu 《Polymer》1982,23(12):1775-1779
Intrinsic viscosity measurements were carried out on five well characterized fractions of poly(ethylene oxide) in aqueous solutions at 24.9°, 34.9°, and 45.5°C. The Stockmayer-Fixman extrapolation was applied to the data: it yields the unperturbed dimensions K0 of the chain. The unperturbed root-mean-square end-to-end distance R?2120 calculated for the polymer fractions in water indicate that the polymer molecules are expanded in this solvent as the temperature is raised. The temperature coefficient of unperturbed dimension, d InR?20dt= 0.024 K?1, calculated for poly(ethylene oxide) in water using the present data is about 100 times higher than the literature values of 0.23 (±0.02) × 10?3 K?1 and 0.2 (±0.2) × 10?3 K?1, respectively, obtained from force-temperature (‘thermoelastic’) measurements on elongated networks of the polymer in the amorphouse state and form viscosity measurements on this polymer in benzene. A value of θ=108.3°C was obtained from the temperature dependence of the interaction parameter B in the Stockmayer-Fixman equation.  相似文献   

15.
T.A. King  A. Knox  J.D.G. McAdam 《Polymer》1973,14(7):293-296
The diffusion of linear polystyrene under non-theta conditions in butan-2-one has been studied by Rayleigh light scattered linewidth measurements for the molecular weight range of 2.08 × 106 to 8.7 × 106 and as a function of concentration. By extrapolation of diffusion coefficient values to zero concentration we find that D0 = 5.5 × 10?4M??0.561wcm2s?1. The first order concentration dependence kdc changes sign as the molecular weight increases, kd being fairly small and negative at low molecular weights and increasingly positive above M?w?230 000.  相似文献   

16.
The isothermal crystallization of poly(ethylene-terephthalate) (PETP) fractions, from the melt, was investigated using differential scanning calorimetry (d.s.c.). The molecular weight range of the fractions was from 5300–11750. Crystallization temperatures were from 498–513 K. The dependence of molecular weight and undercooling on several crystallization parameters has been observed. Either maxima or minima appear at a molecular weight of about 9000, depending on the crystallization temperature. The activation energy values point to the possibility of different mechanisms of crystallization according to the chain length. A folded chain process for the higher M?n chains and an extended chain mechanism for the lower M?n chains. The values of the Avrami equation exponent n vary from 2–4 depending on the crystallization temperature; non-integer values are indicative of heterogeneous nucleation. The rate constant K depends on Tc and M?n, showing maxima related to the Tc used. The plot of log K either vs. (ΔT)?1 and (ΔT)?2 or TmT(ΔT) and T2mT(ΔT)2 is linear in every case.  相似文献   

17.
Extremely high molecular weight polystyrenes with a M?w in the range 10.8 × 106 to 2.2 × 107 were prepared by emulsion polymerization initiated with a heterogeneous initiator at 30°C, which has a ‘living character’. Samples of polystyrene were characterized by light scattering and viscometry in toluene and benzene at 25°C, and in θ-solvent cyclohexane at 34.8°C. Also determined were the relationships of mean-square radius of gyration 〈s2〉 (m2) and the second virial coefficient A2 (m3 mol kg?2) on the molecular weight, which for toluene and benzene are described in equations: Toluene (25°C) 〈s2〉=1.59 × 10?23M?w1.23; A2=4.79 × 10?3M?w?0.63; Benzene (25°C) 〈s2〉=1.23 × 10?22M?w1.20; A2=2.59 × 10?3M?w?0.59. The parameters in the Mark-Houwink-Sakurada equation were established, for extremely high molecular weight polystyrene in toluene and in benzene, at 25°C into the form giving for [η] (m3kg?1): [η] = 8.52 × 10?5M?w0.61; [η] = 1.47 × 10?4M?w0.56. The mentioned relations, as well as the obtained values of Flory parameter ?0 and of ratio [η]M?w0.5 were compared with solution properties of high molecular weight polystyrene with narrow molecular weight distribution prepared by anionic polymerization by Fukuda et al.  相似文献   

18.
The chemical homogeneity of a series of copolymers obtained by nucleophilic substitution of organolithium reagents RCH2Li [R = S(C6H5), SOCH3, SO2CH3 and SO2N(CH3)2] on a high molecular weight poly(methyl methacrylate) (PMMA, DPn = 700) has been studied by different methods, over a wide range of substitution degrees (0.14 ? DSm ? 0.76). ‘Cross’ fractionation is much more efficient than ‘one direction’ fractionation, and it allows the determination of \?gs2 variance values as low as 2 × 10?4 with sufficient accuracy. The light scattering method is far less sensitive in this range of low compositional polydispersity, even for a suitable system. The fairly high chemical homogeneity observed for all the copolymers, prepared either in homogeneous or heterogeneous solution, may be correlated with both the high molecular weight of the PMMA precursor and the autoretarded kinetics of the substitution process.  相似文献   

19.
Laser light scattering including angular dependence of total integrated scattered intensity and of the spectral distribution has been used to characterize five samples of poly(1,4-phenylene terephthalamide), PPTA (commercially known as Kevlar), of different molecular weights in 96% sulphuric acid and 0.1 NK2SO4. The data are supplemented by intrinsic viscosity measurements used to detect the possible effects of association, by differential refractometry providing a measure of the refractive index increments in mixed solvents (H2O, H2SO4 and K2SO4) and by spectrophotometry for the extinction coefficient needed in the correction of attenuation in light scattering studies. The results show 〈DZ = 2.11 × 10?5M?W?0.75cm?2s?1 in reasonable agreement with an average of many of the published intrinsic viscosity data obeying [η] = 1.09 × 10?3 Mw1.25 ml g?1 and w expressed in g mol?1.  相似文献   

20.
Toshio Shimizu 《Polymer》1981,22(2):231-234
The light-scattering, osmotic pressure and viscosity data reported in the preceding paper are used to estimate the characteristic ratios and the persistence lengths of poly(maleic anhydride-co-ethyl vinyl ether) (MAn) and poly(maleic acid-co-ethyl vinyl ether) (H-MA) in organic solvents and sodium salt of H-MA (Na-MA) in NaCl aqueous solutions. The characteristics ratios obtained by Stockmayer-Fixman's extrapolation method are as follows: 6.5 for MAn in organic solvents, 9.6 for H-MA in organic solvents, and 10.8 for Na-MA in NaCl aqueous solutions. The persistence lengths of these copolymers were also calculated by applying the wormlike chain model to viscosity data; ~9 A? for MAn in organic solvents and ~12 A? for H-MA in organic solvents. For Na-MA in NaCl aqueous solutions, the persistence lengths vary in proportion to the inverse root of the ionic strength Cs (i.e., C?12s). The persistence length in the absence of electrostatic interaction (i.e., Cs←∞) is about 10 Å which corresponds to the values of H-MA in organic solvents.  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