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.
At 371.5 K, which is the θ-temperature for polystyrene (PS) in 3-methyl cyclohexanol (MC), intrinsic viscosities [η] have been measured for PS samples of different relative molar mass M in mixtures of MC with a thermodynamically good solvent 1,2,3,4-tetrahydronaphthalene over the whole range of solvent composition. Eleven graphical procedures have been utilised and assessed in deriving the unperturbed polymer dimensions expressed as Kθ (in the relation [η] = KθM12α3 where α is the expansion factor). For those procedures concluded to be the most reliable, there was no influence of binary solvent composition: the value of Kθ = 78 (±1) × 10?3 dm3 kg?1 was the same as that obtained directly under θ-conditions.  相似文献   

2.
θ-Conditions, the temperature coefficient of unperturbed dimensions of the macromolecules and the thermodynamic interaction parameters ψ and κ were determined for solutions of poly(methyl methacrylate) in 2-alkoxyethanols (methoxy, ethoxy and butoxy). The results for this series of solvents fit the data reported for other solvents and dln r02/dT = 2.6 × 10?3K?1. The dependence of parameters ψ and κ exhibited deviations from the theoretical dependence, mainly near the limiting value ψ = 0.5.  相似文献   

3.
We propose the following empirical relationship between the intrinsic viscosity of a polymer and its molecular weight M.
{[η]?[η]θ/[η][η]θ=?Δρ+A′M12
[η] and [η]0 are the intrinsic viscosities in a good solvent and in θ conditions respectively. Δ?, and A′ are constants characteristic of a system polymer-solvent. This relationship is valid for PS and PMMA in various good solvents and for a range of molecular weight from 3000 to 250,000. It is in this range that the Mark-Houwink-Sakurada equation is least applicable.  相似文献   

4.
For the polyamide Trogamid TR, good as well as θ-solvents are available. Determinations of molecular weights by ultracentrifugation, light scattering, osmometry, viscometry and gelchromatography (GPC) are reported. For DMF at 25° the relation
[η] = 0.02737 M0.706w cm3 g?1
was established, [η] = f(T) shows a maximum between ? 70° and + 120°. θ-temperatures are 142° for aniline and 62° for pyrroline, the latter having a negative temperature gradient. The unperturbed dimensions are calculated from the viscosity in θ-solvents and in DMF; for the latter, the Stockmayer-Fixman extrapolation was used. Molecular dimensions proved to be small in comparison with those of similar polymers. This effect is due to the three methyl side-groups for each chain unit sterically preventing a more stretched conformation.  相似文献   

5.
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   

6.
The coil dimensions of polystyrene in various solvents were investigated by light-scattering measurements at temperatures between 20° and 90° and under pressures of up to 2000 bar. The measurements were made at a constant scattering angle over a wide range of the scattering parameter (sin2θ/2)/λ2 by varying the wavelength from 254 up to 827 nm. The radius of gyration (r?) for polystyrene in all systems decreases with increasing pressure. The decrease of r? for these systems becomes less with greater distance from theta-temperature. In contrast with an earlier hypothesis [1], the experiments show that r? at theta-temperature (r?θ), and thus the unperturbed dimensions, are independent of pressure within ±5%, whilst theta-temperature rises with increasing pressure. The theories of excluded volume are applied to describe the characteristics of coil dimensions; the experimental results fitted best the theory of Yamakawa-Tanaka and that of Ptytsin. Dividing the excluded volume of a segment of the polymer chain into a temperature-dependent and a temperature-independent part, the measurements show an increase of the temperature-independent part with rising pressure. The flexibility of the polymer chain in polystyrene solutions rises with increasing pressure, whilst the range of the interaction forces narrows.  相似文献   

7.
A neutron diffraction study has been carried out on Sr0.5La1.5Li0.5Fe0.5O4 of K2NiF4-type derived structure and it has shown that iron in the tetravalent state has a high spin configuration (t32ge1g) and that the material has some stacking defects. At room temperature this compound shows an ordering between iron and lithium atoms leading to a nuclear cell a0√2, a0√2, c0 (a0 and c0 are the parameters of the K2NiF4-type cell). At low temperature (T < TN2) the magnetic structure can be described as antiferromagnetic, corresponding likely to a colinear pattern with a propagation vector of 0.5 (a0) along the [110] axis. At higher temperature (TN2 < T < TN) helimagnetic structure is consistent with a propagation vector of 0.47 (a0) [110].  相似文献   

8.
The theta temperature for the system poly(o-chlorostyrene)-methyl ethyl ketone has been determined as 24·5°. The samples used in the determination were prepared by radical polymerization. The dependence of intrinsic viscosity on molecular weight has been measured in methyl ethyl ketone at 24·5° and found to be ηθ = 4·68 × 10?4MwM12. The ratio 〈s=2〉/M was found, by light scattering, to be 5·60 × 10?18 cm2. Analysis of the solution properties indicates that the Kurata-Yamakawa theory is valid in the vicinity of the Flory temperature (UCST).  相似文献   

9.
Poly[2-methoxy-4,6-di(p,p′-isopropylidene diphenyloxy)-s-triazine] has been synthesized by interfacial polycondensation of 2-methoxy-4,6-dichloro-s-triazine with bisphenol A. The conditions have been optimized to prepare a sample of both high solubility and molecular weight. The polymer has been fractionated by fractional precipitation and the fractions characterized by viscometry, osmometry and gel permeation chromatography. Mark-Houwink-Kuhn-Sakurada (MHKS) expressions have been developed for three solvents. The values of unperturbed dimensions (<R2>0/M)12, solvent-polymer interaction parameter (B and χ1) and conformational parameter (σ) have been computed by applying the two-parameter theories of excluded volume developed by Fixman, Kurata, Stockmayer and Roig and by Flory, Fox and Schaefgen.  相似文献   

10.
Translational diffusion, velocity sedimentation and viscometry of polyamidobenzimidazole (PABI) solutions in the range of M = (1–61) · 103 have been investigated in N,N-dimethylacetamide (DMA) and 98% H2SO4. The dependences of D0, S0 and [η] on M were obtained. Tsvetkov-Klenin's hydrodynamic invariant was found to be A0 = 3.55 · 10?10erg deg?1mol?13. The equilibrium rigidity of PABI molecules was characterized by the length of the Kuhn segment A = 250 ± 100 A?. The chain diameter was 7 ± 4 A?. The values of A in 98% H2SO4 and in an aprotic solvent, DMA, were virtually identical, implying that the rigid-chain conformation of PABI molecules in 98% H2SO4 is due to their geometrical structure rather than to the protonization of amide bonds. The significance of the latter evidently increases in PABI solutions in 100% H2SO4 in which A is 1.5 times as high. The decrease in rigidity of PABI as compared to that of poly-p-phenylene terephthalamide (A = 400 ± 100 A?) is in reasonable agreement with the presence of imidazole rings in PABI molecules. The presence of these rings results in kinks in the PABI chain with angles of about 30° and hence, in the depature from parallelism of rotating bonds.  相似文献   

11.
A reassessment of the dilute solution behaviour of amylose acetate in nitromethane has been made, using linear fractions encompassing the low molecular weight region. The temperature dependence of the limiting viscosity number [η] has been measured and is found to be negative. Mark-Houwink relations have been established at 298° K and 303° K and the unperturbed dimensions have been calculated giving (〈r20M)12 = 589 × 10?11cm at 298°K. The similarities and differences between our data and those of Banks and Greenwood are discussed. The results support the idea that amylose acetate is a stiff coil in good solvents.  相似文献   

12.
For a number of fractions and unfractionated samples of polylaurolactam, molecular weights (Mw = 1 × 104?12·5 × 104) were measured by the light-scattering method in a mixed solvent of m-cresol with 60 vol. % 2,2,3,3-tetrafluoropropanol; intrinsic viscosities were determined in m-cresol and 96% H2SO4, and the constants of the Mark-Houwink equation were calculated. The calculated values of the characteristic ratio of unperturbed dimensions (virtually identical for m-cresol and 96% H2SO4) were compared with the respective values for polypyrrolidone and polycaprolactam. It was found that the higher frequency of the CONH-groups reduces the rigidity of the polyamide chain.  相似文献   

13.
Reactions between MX(PPh3)2(η-C5H5) (M = Ru, X = Cl; M = Os, X = Br) and 2-CH2CHC6H4PPh2 afford MX(η2-CH2CHC6H4PPh2)(η-C5H5); the Os complex is obtained in two isomeric forms. The X-ray structure of the major isomer shows the CC double bond (OsC, 2.214, 2.195 Å; CC, 1.57 Å) is almost coplanar with the OsBr vector, with the terminal C cis to Br; the minor isomer is assumed to have the alternative, more sterically congested conformation, with the β-C cis to Br. The chlororuthenium complex reacts with NaOMe/MeOH to give the corresponding hydrido complex, which also exists as two isomers in solution; reaction of this complex with CS2 gives the expected dithio acid derivative Ru(S2CCHMeC6H4PPh2)(η-C5H5), together with small amounts of a complex assumed to be Ru[S2C(CH2)2C6H4PPh2](η-C5H5). The X-ray structure of the major product reveals an unusual η3-S2C mode of coordination of the dithio acid fragment (RuS, 2.418, 2.426(1) Å; RuC 2.175(4) Å). Crystals of OsBr(η2-CH2CHC6H4P)Ph2)( η-C5H5) are monoclinic, space group P21/n, with a 12.696(2), b 21.719(6), c 15.929(3) Å, β 79.77(2)°, Z = 8; 2867 data (I > 2.5σ(I)) were refined to R = 0.040, Rw = 0.044. Crystals of Ru(η3-S2CCHMeC6H4PPh2)(η-C5H5) are orthorhombic, space group Pbca, with a 8.921(2), b 15.982(9), c 32.216(5) Å, Z = 8; 1685 data (I > 2.5σ(I)) were refined to R = 0.027, Rw = 0.030.  相似文献   

14.
15.
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   

16.
An ultra-high molecular weight and narrow distribution polystyrene (Mw = 7.3 × 106, Mw/Mn = 1.13) was dissolved in a wide range of solvents. Potential degradation by freezing was studied as a function of solvent type, concentration, cooling rate and number of freezing cycles. Cryogenic experiments were conducted in dioxane, tetrahydrofuran, benzene, dichloroethane, cyclohexanone, p-xylene, methyl methacrylate and styrene. The extent of degradation did not relate to a single solvent parameter, but there seemed to be a tendency towards a limited degradation in solvents with low melting points and/or solubility parameters greatly different from that of polystyrene. A low polymer concentration as well as a high cooling rate promoted chain scission, the latter parameter being the most important. In cyclohexanone and p-xylene, linear relationships were observed between the number of scission per molecule and the number of freezing cycles at high polymer concentrations and at high cooling rates. At lower concentrations and slower cooling, the relationships were non-linear suggesting a different degradation mechanism. The most extensive change in molecular weight distribution was observed on freezing in styrene. After 45 freezing cycles, an Mw of only 2.3 × 106 was observed. The results indicate that chain scission occurred together with polymerization and combination reactions. Freezing of suitable solutions of ultra-high molecular weight polymers can thus be used as a new way of initiating polymerizations by cooling rather than heating.  相似文献   

17.
(η-Cyclopentadienyl)(triphenylphosphine)cobaltacyclopentadienes having an electron withdrawing substituent on the cyclopentadienyl ring, (η-C5H4R)(PPh3)(CoCHCHCH) (1b: R = COOMe; 1c: R = COMe), were prepared in reasonable yields by treatment of a solution of (η-C5H4R)(PPh3)2Co with acetylene. A non-substituted cyclopentadienyl analog (1a: R = H) was also isolated in low yield according to a similar procedure. Novel dinuclear complexes were also formed as by-products and the structure of (η-C5H4R)Co(PPh2C6H4)(μ-CMe)Co(η-C5H4R) (2b: R = COOMe), having a μ23-benzyl moiety, was determined by an X-ray crystallographic analysis. The X-ray analyses of 1a and 1b were also carried out. Crystals of 1a are monoclinic, space group Pa, a 8.529(3), b 16.010(6), c 8.028(4) Å, β 100.31(3)°, Z = 2; crystals of 1b are monoclinic, space group P21/a, a 8.327(2), b 36.468(7), c 8.021(1) Å, β 98.75(2)°, Z = 4; and crystals of 2b are monoclinic, space group P21/c, a 10.681(2), b 30.722(7), c 8.912(1) Å, β 93.55(1)°, Z = 4. They have been refined to R = 0.034, 0.047 and 0.050, respectively.  相似文献   

18.
The study of K2NiF4 and perovskite structure type by the “method of invariants” leads to the relationship: (A-X)9 212 ? (A-X)12 = constant, where (A-X)9 and (A-X)12 are the invariant values associated with cation A in coordination number 9 and 12. In the case where A = K+ and X = F?, we propose the relationship:
(K+?F)R = 2.832 R111.4
where R is the coordination number.  相似文献   

19.
The method of reduced variables or superposition is applied to investigate the first normal stress difference (N1)-temperature dependence over the shear rate (γ) range 10–1000 s?1 for polystyrene (PS) and high-impact polystyrene (HIPS) melts, at temperatures of 180 C and above. These conditions are similar to those for industrial polymer processing. For PS, the first normal stress differences are obtained using Tanner's equation, leading to good agreement with values obtained by other authors and methods. For HIPS we have not found data in the literature for N1 at shear rates above 10s?1. In our case N1 was obtained by measurements of entrance pressure losses. Correlations, based on master curves, are found for first normal stress difference in terms of shear rate and temperature. All samples follow the power law equation N1 = kγ?m, with values of m ranging from 0.50 to 0.64.  相似文献   

20.
The radius of gyration of polystyrene molecules in the bulk polymer has been measured by low angle neutron scattering from a dilute solid solution of poly proto-styrene in a matrix of polydeutero-styrene. The radius of gyration is 90 A? ± 5 percent (Ma = 97,200), which agrees within experimental error with the unperturbed dimensions of the polymer chain (? 84 Å) as determined from the solution properties of the polymer.  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