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.
The cationic polymerizations of methyl-, 2-chloroethyl-, ethyl-, cyclohexyl- and t-butyl- vinyl ethers initiated by cycloheptatrienyl hexachloroantimonate in methylene chloride solutions have been studied in detail. Reaction rates were measured by an adiabatic calorimetric technique and rate constants for propagation of each of the monomers, kp (obs), were determined by appropriate kinetic analysis of the experimental curves. The results obtained are discussed in terms of current theories regarding ion pair/free ion equilibria in non-aqueous solvents. Although ethyl-, cyclohexyl- and t-butyl- vinyl ethers behave very similarly to isobutyl vinyl ether, and their reactivities are comparable [kp (obs) ~ 3 × 103M?1sec?1 at 0°C] both methyl- and 2-chloroethyl- vinyl ethers show markedly different characteristics to the others, and in particular exhibit a reactivity approximately one order of magnitude less [kp (obs) ~ 2 × 102M?1sec?1 at 0°C]. These variations in reactivity are discussed in terms of preferred monomer conformations, and the resulting differences in activation energy which are likely to arise when such conformers are approached by an electrophile.  相似文献   

2.
B. Nyström  J. Roots  R. Bergman 《Polymer》1979,20(2):157-161
Sedimentation velocity measurements on polystyrene (M = 110 000) in cyclopentane over an extended concentration region and from 5°C (close to the upper critical solution temperature) to 40°C are reported. The concentration dependence parameter (ks)w increases from 2 to 5°C to 27 at 40°C. For all temperatures except 5°C, s0s vs. w[η]w shows an upward curvature at w[η]w ≈ 1; at 5°C, on the other hand, s0s is independent of concentration over the region considered. Furthermore, measurements have also been performed at 20°C (θ-conditions) over a large concentration interval for the molecular weights M = 20 400, 390 000 and 950 000. The parameters s0 and (ks)w were both found to be proportional to M?1/2w. In the ‘hydrodynamically normalized’ plot s0s vs. w[η]w the sedimentation behaviour can approximately be represented by a single curve for all the molecular weights.  相似文献   

3.
The pressure dependence of the upper critical solution temperature (dTdp)c in the polystyrene-cyclohexane system has been measured over the pressure range of 1 to 50 atm. The value of (dTdp)c determined over the molecular weight (Mw) range of 3.7 × 104 to ~145 × 104 greatly depends on the molecular weight of polystyrene. The value of (dTdp)c for a polystyrene solution of low molecular weight (Mw = 3.7 × 104) is positive (3.14 × 10?3 degree atm?1), while the values are negative (?0.52 × 10?3~?5.64 × 10?3 degree atm?) for solutions of polystyrene over the high molecular weight range of 11 × 104 to ~145 × 104. The Patterson-Delmas theory of the corresponding state and the newer Flory theory have been used to explain this behaviour.  相似文献   

4.
D.J. Blundell  B.N. Osborn 《Polymer》1983,24(8):953-958
The morphology and related properties are described for the aromatic thermoplastic poly(aryl-ether-ether-ketone) (PEEK) [C6H4OC6H4OC6H4CO]n. Topics covered include crystallinity, crystallization and melting behaviour, Iamellar thickness and spherulitic structure. The data are used to derive the following material parameters T1m = 395°C, σe = 49 erg cm?2, σs = 38 erg cm? and ΔHF = 130 kJ kg?1. PEEK is closely analogous to poly(ethylene terephthalate) in its crystallization behaviour except that the main transitions occur about 75°C higher.  相似文献   

5.
C Price  G Allen  N Yoshimura 《Polymer》1975,16(4):261-264
Thermomechanical heat of torsional deformation measurements have been made on crosslinked cis-polybutadiene by means of a Calvet microcalorimeter operated at 30°C. When corrected for volume changes utilizing the Gaussian statistical theory of elasticity, the data gave a value for the relative energy contribution to the torsional couple, MeM, of 0.14 ± 0.02. Measurements were also made on a sample subjected to simple tensile deformations. The relative energy contribution to the tensile force (fef) was found to agree within experimental error with the value obtained for MeM, and the two results gave an average value for din 〈r20dT of 4.1 × 10?4 K?1.  相似文献   

6.
Wyn Brown  Peter Stilbs 《Polymer》1983,24(2):188-192
Transport in ternary polymer1, polymer2, solvent systems has been investigated using an n.m.r. spin-echo technique. The dependence of the self-diffusion coefficient of poly(ethylene oxide) polymers on the concentration and molecular size of dextran in aqueous solution has been measured. Monodisperse poly(ethylene oxide) fractions (M?w=7.3×104, 2.8·105 and 1.2·106) and dextrans (M?w=2·104, 1·105 and 5·105) have been employed over a range of concentration up to the miscibility limit in each system. It is found that when the molecular size of the diffusant is commensurate with or exceeds that of the matrix polymer, a relationship of the form: (DD0)PEO=exp?k(C[η]) is applicable, where C[η] refers to the dextran component and is considered to describe the extent of coil overlap in concentrated solution. (DD0) is independent of the molecular size of the poly(ethylene oxide), at least in the range studied (Mw<300 000).  相似文献   

7.
G.B. McKenna  K.L. Ngai  D.J. Plazek 《Polymer》1985,26(11):1651-1653
Within the context of a generalized coupling model we can support the hypothesis that, while the mode of relaxation for self diffusion (D) and shear flow (η) are the same, the entanglement interactions are different. We assume that there are two distinct coupling parameters nD and nη for self diffusion and shear flow respectively. The model predicts the molecular weight and temperature dependences to be scaled by the relevant coupling parameters as:
η∝[M2exp(Ea/kT)]1(1?nη)and D∝M[M2exp(Ea/kT)]?1(1?nD)
for melts with Arrhenius temperature dependences. We have found that nn=0.43 and 0.42 for polyethylene (PE) and hydrogenated polybutadiene (HPB) which scale η as M3.5 and M3.4. Also the apparent flow activation energies E1a of 6.35 kcal mole?1 for PE and 7.2 kcal mol?1 for HPB scale to primitive activation energies Ea of 3.6 and 4.2 kcal mole?1 for PE and HPB respectively. On the other hand the M?2 dependence of D results in nD=1/3. Then the reported activation energies for self-diffusion in PE and HPB of 5.49 and 6.2 kcal mole?1 scale to primitive activation energies of 3.7 and 4.1 kcal mole?1, respectively.  相似文献   

8.
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   

9.
The synthesis and characterization of methacrylate-ended macromers (M?n 500 to 10 000) and their copolymerization with styrene (M2) is described. The experimental errors in the values of the reactivity ratios r1 render them meaningless. Values of r2 can be determined with more precision and increase from 1.06 to 1.55 as the molecular weight of the macromer increases. This behaviour is due to steric effects, not diffusion-controlled propagation. It is shown that the assumptions that 1 > r1[M1][M2] and r2 >[M1][M2] are only valid for macromers of M?n > ca. 10 000.  相似文献   

10.
Polymerization, and copolymerization with styrene, of m,p-chloromethylstyrene have been carried out at 75°C, in chlorobenzene and in the presence of AIBN ([AIBN] ? 6 × 10?2, and 12 × 10?2m, respectively). The polymer molecular weights, determined by g.p.c., are: M?w = 8670, M?n = 5860, and M?w/-Mn = 1.48 for the homopolymer, poly(m,p-chloromethylstyrene), (1a); and M?w = 8805, M?n = 5144, and M?w/-Mn = 1.71 for the copolymer, copoly(m,p-chloromethylstyrene-styrene), (2a). A series of phosphine derivatives of both 1a and 2a are prepared by the reaction of the polymers with either chlorodiphenylphosphine/lithium, or diphenylphosphine/potassium tert. butoxide. A number of other potentially electroreactive derivatives of 2a are obtained by reacting the polymer with 2-aminoanthraquinone, 3-N-methylamino-propionitrile, or 2-(2-aminoethyl) pyridine. The phosphinated polymers are reacted with bis-benzonitrilepalladium-(II) chloride to obtain a series of polymer-palladium(II) complexes containing 8.5–12.9% palladium. Similarly, reaction of the last-named bidentate polymeric ligand with cupric acetylacetonate, or cupric sulphate pentahydrate, produces polymer-copper(II) complexes having 5.8, or 3.3% copper, respectively. The inter/intra-chain nature of some of the side reactions during the derivatization of the chloromethylated polymers, and that of the complex formation between transition metal centres and macromolecular ligands, are briefly discussed in view of the experimental results.  相似文献   

11.
Yasuhiko Onishi 《Polymer》1980,21(7):819-824
Effects of the molecular weight of dextran on its graft copolymerization with methyl methacrylate (MMA), initiated by ceric ammonium nitrate (CAN), have been investigated. The results indicate that grafting (%), graft polymerization (%) (ψ), the overall rate constant (k′) for consumption of Ce4+, and branch PMMA were influenced significantly by the molecular weight of the backbone polymer dextran. The number of branch PMMA chains per dextran molecule was 0.05 ~ 0.30 for M?w 9000 dextran (D1), 0.35 ~ 0.55 for M?w 61 000 (D2), and 0.8 ~ 1.6 for M?w 196 000 (D3), respectively. The relationship between the rate of graft polymerization and M?w (the weight-average molecular weight of dextran) was expressed by the equation: Rpg = ?AlogM?w + B. Another linear relationship was obtained between In (100 ? ψ) and reaction time (t) for both D1 and D2 samples or In t for D3. Detailed kinetic analysis has been made on the basis of the latter relationship. Mechanical properties were also studied on the moulded sample plates of these copolymers.  相似文献   

12.
K. Takaya  H. Tatsuta  N. Ise 《Polymer》1974,15(10):631-634
Living anionic polymerization of styrene was kinetically investigated in triglyme-benzene mixtures. At low concentrations of triglyme the overall propagation rate constant, kp, was much larger than at the same concentration of monoglyme (DME) in DME-benzene mixtures. The Szwarc-Schulz plot did not have negative slopes for lithium and sodium salts at triglyme contents of 5~20vol%, and no contribution of free anions to the propagation was observed for the sodium salt. The sodium ion pair was more highly reactive than the lithium ion pair; thus at 25°C, the ion pair rate constant, kp, for the lithium salt was 43, 102, 135 and 165 M?1sec?1 at triglyme concentrations of 5, 10, 15, and 20%, respectively, while that for the sodium salt was 410, 920, and 1460 M?1sec?1 in 5, 10, and 15% triglyme, respectively. The dissociation constant, K, for the lithium salt was 2·4×10?11, 1·9×10?10 and 1·3×10?9 M in 10, 15, and 20% triglyme, respectively and the free ion rate constant, kp, was 2~2·5×104 M?1sec?1 for the lithium salt.  相似文献   

13.
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   

14.
N. Kuwahara  M. Nakata  M. Kaneko 《Polymer》1973,14(9):415-419
Cloud-point curves for solutions of five polystyrene samples, including three well-fractionated polystyrenes, in cyclohexane have been examined near their critical points. Even for a solution of polystyrene characterized by MwMn<1.03, the critical point determined by the phase-volume method is generally situated on the right hand branch of the cloud-point curve. The precipitation threshold concentration is appreciably lower than the critical concentration, while the threshold temperature slightly deviates from the critical temperature. The agreement of the precipitation threshold point with the critical point has been found for a solution of polystyrene characterized by Mw=20×104 and MwMn<1.02 in cyclohexane. The η(φ) function derived from critical miscibility data is expressed by χ(φ) = 0.2798+67.50T+0.3070φ+0.2589φ2, which yields θ of 33.2°C and ψ1 of 0.22.  相似文献   

15.
John D. Hoffman 《Polymer》1982,23(5):656-670
The theory of polymer crystallization with chain folding is extended to include the effect of reptation in the melt on the rates of crystallization GI and GII in régimes I and II. The result is that the pre-exponential factors for GI and GII contain a factor 1n, Where n is the number of monomer units in the pendant chain being reeled onto the substrate by the force of crystallization; n is proportional to the molecular weight. The predicted fall in growth rate with increasing molecular weight is found experimentally in nine polyethylene fractions Mz=2.65 × 104 to Mz=2.04 × 105, corresponding to nz=1.90 × 103 to 1.45 × 104. The data on these fractions are analysed to find the reptation or ‘reeling’ rate r and the substrate completion rate g. The values gnuc~0.5/nz cm s?1 and rnuc~21/nz cm s?1 at 400K are obtained from the data in conjunction with nucleation theory adapted to account for reptation assuming a substantial degree of regular folding. These results are consistent with a melting point in the range of ~142° to ~145°C. (The analysis using T°m(∞)=145°C gives values of such quantities as σ σe and α that are quite similar to those deduced in earlier studies.) An estimate of g (denoted gexpt) that is independent of the molecular details of nucleation theory gives gexpt~0.4/nz cm s?1 and r~17/nz cm s?1 at 400K. Calculations of the reptation rate from r1,2 = (force of crystallization ÷ friction coefficient for reptation in melt), where the friction coefficient is determined from diffusion data on polyethylene melts, leads to r1,2~17/nz to 34/nz cm s?1 at at 400K, or g1,2~0.4/nz to 0.8/nz cm s?1. The conclusion is that the reptation rate characteristic of the melt is fast enough to allow a significant degree of adjacent re-entry or ‘regular’ folding during substrate completion at the temperature cited, and that the substrate completion process is governed jointly by the activation energy for reptation Q1D and the work of chain folding q. The nucleation theory and the friction coefficient theory approaches are compared, and the formulations found to be essentially equivalent; the ‘reeling’ rate r is found to be proportional to (1n)A0(Δf)v0exp[?(Q1D+q)RT], where v0 is a frequency factor, and A0(Δf) is the force of crystallization on the pendant chain. The data analysis on the fractions confirms the detailed applicability of régime theory. The growth rate theory presented allows the possibility that the growth front may be microfaceted in régime I.  相似文献   

16.
The soluble V(acac)3-Al(i-C4H9)2Cl system initiated living polymerization of propene at ?78°C affording monodisperse polymers (M?wM?n = 1.15 ± 0.10). A kinetic study (of the living polymerization) was carried out to evaluate the rate coefficients for propagation. The equilibrium constant KM for a propene monomer coordinated to an active vanadium and the rate constant kp for a subsequent insertion of coordinated monomer into a living polymer chain were determined and compared with the values for the polymerization of propene with other soluble vanadium-based catalyst systems. The relation between KM and kp revealed that a strong interaction between vanadium and propene is unfavourable for the insertion of the coordinated propene into a living polymer chain. The mechanism of an initiation reaction involving alkylation and complexing of V(acac)3 with Al(i-C4H9)2Cl has been proposed.  相似文献   

17.
Light scattering measurements for two samples of polystyrene (I, Mw = 2.15 × 105; II, Mw = 2.5 × 106) were performed in the iso-refractive mixed solvent dimethoxymethane-diethyl ether. For sample I the temperature dependence of the second osmotic virial coefficient A2 was determined for three constant compositions of the mixed solvent. In the range ?30° to +25°C the three curves run practically parallel and exhibit a maximum at approximately ?10°C. For the volume fraction of 0.7 diethyl ether in the mixed solvent, an endothermal theta-temperature θ+ was found at ?27.0° ± 1.5°C and θ?, the exothermal theta-temperature, at ?5.0 ± 1°C. The investigation of sample II in the abovementioned solvent confirmed the observed θ?-temperature and displayed a higher exothermicity compared with I. Similarly to the temperature variation of A2, the chain dimensions of II, determined from the angular dependence of the scattered light, run through a maximum. The unperturbed dimensions in the mixed solvent are found to be: rw = 448 ± 5 A? at θ+ = ?27°C and rw = 443 ± 5 A? at θ? = ?5°C, as compared with rw = 420 ± 10 A? at θ+ = +33°C in cyclohexene. The inter-relation of the chain expansion coefficient and A2 is quantitatively described by the Zimm-Stockmayer-Fixman equation over the entire range of heats of dilution.  相似文献   

18.
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   

19.
The interactions of bis-2-(2-pyridylazo)-1-naphthol Co(III), [Co(III) (αPAN)2+], with five kinds of synthetic polyelectrolytes have been studied by spectrophotometric and transient electric dichroism measurements. The polyelectrolytes were: poly(styrene sulphonic acid) (PSS); poly(acrylic acid) (PAA); poly(l-glutamic acid) (PLG); poly(Ne,Ne-dicarboxylmethyl-l-lysine (PDCML); and poly(l-lysine) (PLL). The equilibrium constant of the reaction:
with P = polyelectrolyte residue was determined spectrophotometrically: K1 is > 107 M?1 (PSS); (1.4 ± 0.2) × 104 M?1 (PAA); (4.0 ± 0.4) × 103 M?1 (PLG); (1.4 ± 0.2) × 106 M?1 (PDCML); and < 102 M?1 (PLL) at pH 6–8. From transient electric dichroism, the angle (ψ) between the αPAN plane and the polymer axis was determined to be 65° (PSS); 52° (PAA); 55° (PLG); and 52° (PDCML). The large K1 and ψ values for PSS are ascribed to the hydrophobic interaction between the aromatic αPAN ring and the styrene sulphonate residues of PSS. Using stopped-flow electric dichroism measurements, rapid transfer of a bound Co(III) chelate from the PDCML to PSS chains was shown to occur.  相似文献   

20.
The temperature dependence of stress and birefringence for natural rubber vulcanizates under medium and large deformation was measured for the processes of cooling, heating and re-cooling. In order to investigate the relation between the stress and crystal phase, the observed birefringence, Δt, was converted into the crystallinity, Xv, by the following equation:
Xv = Δt?Δna°faΔnc°fc?Δna°fa
where Δn0c, Δn0a, fa and fc are the intrinsic birefringence of the crystal, that of the amorphous phase, the orientation factor of crystallites, and that of amorphous phase, respectively. The fusion of crystallites induced by the thermal crystallization resulted in the increasing contractile force, while the fusion of strain-induced crystallites induced the reduction of contractile force.  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