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 656 毫秒
1.
The bulk viscosities η of over fifty sharp fractions of cyclic and linear poly(dimethyl siloxanes) in the weight-average molecular weight range 500 < M?2 < 25 000 have been measured at 298 K using a cone- and-plate microviscometer. In the Iow molecular weight region M?W < 1000) the η values for the cyclics were found to be at least three times as large as the values for the corresponding chain molecules. By contrast, in the highest molecular weight region (M?W > 16 000), the η values for the cyclics were approximately one-half those for the corresponding linears. Cyclics and linears containing about one hundred skeletal bonds were found to have similar bulk viscosities. The temperature dependence of the bulk viscosities of eighteen of the cyclic and linear fractions were investigated, and the relationship η = A exp(EviscRT) was used to deduce values for the energies of activation for viscous flow Evisc and the constants A.  相似文献   

2.
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   

3.
Copolymerization of an equimolar mixture of m,p-chloromethylstyrene (M1) and styrene (M2) was carried out in chlorobenzene in the presence of AIBN at 80°C. Molecular weight analysis (by g.p.c.) of the resulting polymer samples was performed at various conversions. M?w, M?n, and (M?wM?n) value of 21 300, 13 800 and 1.54 were obtained at 8.9% conversion. At higher conversions, the value of M?w remained effectively constant while M?n decreased to 9200 at ca. 80% conversion, and then increased to 12 000 at about 100% conversion (16 h), and to 13 700 if the polymer solutions were maintained at 80°C for an additional 44 h. These results suggest that, although the termination step initially involves the combination of polymer radicals, at high conversions a large number of very low molecular weight, and unsaturated, polymer molecules are formed possibly by disproportionation involving polymer radicals and primary radicals. The unsaturated polymer molecules are subsequently polymerized by growing polymer radicals towards the end of the polymerization. It was noticed that further reaction occurred after complete depletion of monomer, involving radical attack on the unsaturated polymer molecules. Other reactions including chain transfer to polymer will also be important at high polymer concentrations. A copolymer of M1 and M2 was separated into four fractions on a preparative scale, and molecular weight analysis of the resulting polymer samples provided more evidence of the above interpretation. G.p.c. analysis of several derivatives of a copolymer of M1 and M2 showed that most molecular weights were much lower than that of the starting polymer. These results in some cases may reflect the chemical or dimensional changes introduced into the polymer molecules during derivatization.  相似文献   

4.
The variation of refractive index increment dndc with molecular weight has been studied using solutions of monodisperse polystyrenes (2000<Mw<4×106) in benzene. The results are in good agreement with those obtained by several authors using other systems.We have shown that the linear relationship:
dndc=dndcm+K″Mn
where (dndc)m is the refractive index increment of the repeating unit and K″ a constant, holds only for low molecular weights. With the assumption of volume additivity, we have estimated quantitatively the linear portion of the curve, dndc=f(1M), observed for molecular weights below 20 000 as a function of polymer composition. Beyond this limit, the increase of dndc with molecular weight may be probably related to segment-segment interactions within the coil.  相似文献   

5.
Free and covalently bonded (esterified) nitroxyl radicals experienced in poly(ethylene glycols) (PEG; M?n 200–22 000) at temperatures T >Tg several different isotropic rotational relaxation regions. As a first approximation it was assumed, that in the polyglycols M?n ? 1000 there are at least three rotational relaxation regions: the liquid state (I), the melting state (II) and the solid state (III). The existence of the fourth region, the frozen solid state (IV), was also concluded. The existence of the relaxation region (II) indicated the close interaction between radicals and the crystalline phase. The order of rotational activation energies (Ea) was EIIa >EIIIa >EIa >EIVa (M?n ? 1000). In the polymer melts (I) Ea values of free and bonded radicals first diminished as a consequence of the decrease of the end group effect and they achieved constant high molecular weight values (~15 and 25 kJ respectively). Ea changed in the solid state as a function of M?n principally in the same manner except of the higher numerical values (~40 kJ).Ea of free and covalently bonded radicals in the transition region (II) gained a maximum at M?n 1550 (125 and 170 kJ) and another at M?n > 9500 (130 and 165 kJ) expressing the high degree of order in these polymers in the solid state.The results obtained correlated well with the proton magnetic resonance measurements but they did not correlate with the amorphous dielectric relaxation measurements.It was concluded that the following factors may affect the rotational relaxations of nitroxyl radicals in PEG: the free volume of the polymer, the crystallinity, the chain packing and the end-group effect. The segmental character of the relaxation process was clearly indicated.  相似文献   

6.
By combining static and dynamic properties (Mw, A2, kdRg and Rh) of poly(1,4-phenyleneterephthalamide), PPTA (commercially known as Kevlar), with a detailed analysis of measured time correlation functions at different scattering angles in dilute solution, we have been able to estimate the molecular weight dependence of the radius of gyration, Rg(M), the persistence length ? (≈ 290 A?), and the molecular weight distribution (Mz:Mw:Mn ≈ 6.2:1.8:1) using an unfractionated PPTA sample (Mw = 4.3 × 104 g/mole). Laplace inversion of the time correlation function was accomplished independently by means of two different algorithms: the singular value decomposition technique with discrete multi-exponentials to approximate the normalized characteristic linewidth distribution function G(г) and the method of regularization whereby a linearized smoothing operator was used. The non-intrusive laser light scattering technique permits us to characterize, for the first time, the molecular weight distribution of PPTA which has been difficult to perform by means of other more established methods, such as size exclusion chromatography, because of the corrosive nature of solvents used in preparing PPTA solutions.  相似文献   

7.
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   

8.
J.H. Magill  H.-M. Li 《Polymer》1978,19(4):416-422
The crystallization behaviour of polymer blends or mixtures of the same system has been studied over a wide range of molecular weight and crystallization temperatures. Blends were made by mixing fractionated polymer samples. The spherulitic growth in these mixtures is dependent upon the number-average molecular weight of the system at the shorter chain lengths, but then becomes insensitive to molecular weight values when about 105 to 106 are reached. The growth rate kinetics of mixtures can be described by a kinetic model used for fractionated poly(tetramethyl-p-silphenylene siloxane) (TMPS) polymers. The crystal surface energies deduced from these rate data are molecular weight dependent as are the pre-exponential and transport factors in the rate equation. These parameters are explained in terms of the crystallite morphology. Mixtures (as well as fractions themselves) of all polymer fractions ranging from the monomer to the highest molecular weight (106 approximately) have similar morphological features and form negatively birefringent spherulites. Although molecular weight segregation appears to play an important role in crystallization at comparatively small undercoolings, its influence seems to be minimal at large undercoolings (close to or below the growth rate maxima). Very low molecular weight additives significantly affect the overall crystallization kinetics. Compared to the undiluted sample, mixtures so formed have lower observed melting points and glass transition temperatures. Rates of crystallization are generally facilitated by the diluent with the peak in the growth rate being displaced to lower temperatures. The growth rates for diluted over the undiluted polymer at similar undercoolings are usually larger. At high molecular weights the log of the spherulitic growth rate varies as M?12n, over a considerable range, but at low molecular weight values the rate depends more strongly on Mn approaching a limit of M?1.2n as the monomeric state is approached.  相似文献   

9.
Reverse osmosis data on two different cellulose acetate membranes using seven organic solutes of varying molecular weight have been obtained.A combined viscous-flow and frictional model is presented and used to estimate the maximum retention, the friction between solute and membrane, the distribution coefficient for solute and the pore radius.The calculated values of the maximum retention and distribution coefficient have been compared with the Ferry-Faxen equation. For the more open membrane these are in good agreement. The tighter one, however, shows a greater interaction between solute and membrane than predicted by the Faxen equation.Some data on two-solute systems are presented and shown to give variation in the retention, which can be explained from the convection term.Furthermore, for experiments with dextran the permeate shows a significant reduction in both Mn and Mt  相似文献   

10.
Emulsion polymerization of vinylacetate leads to branched polymers which at high monomer conversions form microgels of the shape and size of the latex particles. Quasielastic light scattering measurements from samples in the pre-gel state give at small q2 a linear angular dependence of Dapp = Гq2 which resembles that of randomly branched chain molecules, where Г is the decay constant of the time correlation function. Extrapolation of Dapp towards zero scattering angle yields the translational diffusion constant Dz. The diffusion constant follows the molecular weight dependence Dz = 9.78 10?5Mw?0.478. The diffusion constant of the microgels, i.e. at molecular weights Mw > 14 106, remains constant because of the finite and constant size of the latex particles. The coefficients kf and kD in the concentration dependence of the frictional and diffusion coefficients are related according to the equation kD = kf ? 2A2Mw ? v? where A2 is the second virial coefficient and v? the partial specific volume of the particle. The coefficient kf is calculated from the experimentally determined quantities kD, A2 and Mw, and the result is compared with the theory by Pyun and Fixman. Accordingly the branched coils in the pre-gel state resemble soft spheres, but the microgels behave more like spheres of some rigidity.  相似文献   

11.
Laminar pool flames of four straight-chain alcohols were studied in uncooled glass and copper burners and water-cooled glass burners. The mass burning rate m? increased with the burner diameter d decreasing as m? ∝ d?n, where n ≈ 1.2 for uncooled glass burners. The values of m? in copper burners were lower by a factor of 1.5–2 than those in glass burners of the same diameter. Vigorous boiling of the subsurface layer of liquid occurred in sufficiently narrow uncooled glass burners, and numerous burning droplets were observed to traverse the flame. The quenching diameter dc was measured. For uncooled copper burners dc increased with increasing molecular weight Mr of the alcohol, from ≈ 2 mm for propanol to ≈ 20 mm for decanol. For uncooled glass burners dc was much lower (1–2 mm) and less dependent on molecular weight. The product of dc and the critical mass burning rate is assumed to have a characteristic value for the stability of diffusion burning, in accordance with the Zel'dovich criterion (Pe) = (Pe)c = const. for flame stability in gas mixtures. However, the stability of diffusion burning increased with increasing heat losses from the burner wall. Two possible interpretations of this paradoxical effect are suggested.  相似文献   

12.
Differential scanning calorimetry (d.s.c.) was used to investigate the thermal behaviour of cyclic and linear poly(dimethylsiloxanes) over the temperature range 103–298 K. Fractions of the polymers studied had number-average molar masses in the range 160 < Mn < 25 500 g mol?1 and heterogeneity indices MwMn < 1.1 in most cases. D.s.c. was applied to measure the glass transition temperatures Tg cold crystallization temperatures Tc and polymer crystalline melting temperatures Tm of the oligomer and polymer fractions. Cyclic siloxanes [(CH3)2SiO]x with number-average numbers of skeletal bonds nn in the range 24 ≦ nn ≦ 79 and linear siloxanes (CH3)SiO[(CH3)2SiO]ySi(CH3)3 with nn in the range 10 ≦ nn ≦ 40 were found not to crystallize. The Tg values of the linear siloxanes were found to be in agreement with values in the literature and they increased with increasing Mn. By contrast, the Tg values of the cyclics were found to decrease with increasing Mn.  相似文献   

13.
Small-angle neutron scattering studies have been made of molten and crystalline polypropylene using samples containing small amounts of deuterated polypropylene in a protonated polypropylene matrix. The specimens were characterized by small- and wide-angle X-ray scattering to determine the d-spacing and the degree of crystallinity χ and by gel permeation chromatography to determine molecular weight, Mw, and molecular weight distribution. The degree of crystallinity was varied from 0.5 to 0.7, the d-spacing from 120 to 250 Å and the molecular weight from 34 000 to 1 540 000. Clustering was not observed. The radius of gyration 〈s2w12 of the tagged molecules was approximately proportional to Mw12 and almost independent of d and χ. In the melt similar values were obtained which are, within experimental uncertainties, the same as in a θ-solution. For 〈s2wk2? 1 the scattering law approaches a k?2 dependence. The results are discussed with reference to the chain-folded model but a fit cannot be obtained over all molecular weights. A simple random coil model fits the neutron scattering data partly but this does not explain the origin of the d-spacing.  相似文献   

14.
Bruno Roland  Johannes Smid 《Polymer》1984,25(8):1166-1172
Binding constants of 1-pyrenebutyrate (PB?, 1-pyrenecarboxylate and 1-pyrenevaleriate to the polysoap-type macromolecule poly(vinylbenzo-18-crown-6) (P18C6) and the poly(vinylbenzoglyme) PVBG (a polystyrene with two CH3O(CH2CH2O)2-substituents at the 3 and 4 position of each benzene ring) were determined spectrophotometrically (λm 342 nm for free PB?, 348 nm for polymer-bound PB?). The binding appears to follow Langmuir adsorption behaviour. Interaction with P18C6 is enhanced on adding crown ether-complexable cations, which converts the neutral polymer into a polycation. The enhanced binding is chiefly caused by lower values of 1n (minimum number of crown monomer units per bound PB? molecule). This decreases at 25°C from 1n = 25, for neutral P18C6, to 6.2 and 2.7 in the presence of 0.01 M CsCl and 0.1 M KCl, respectively. It is argued that the COO? substituent of bound PB? is probably located in the aqueous layer at the polymer-water interphase. Its presence close to the crown ligands enhances the binding of K+ or Cs+ cations to these ligands by forming crown-complexed ion pairs PB?, M+ … P18C6. The typical excimer fluorescence emission peak of PB? is observed when P18C6 in 0.1 M KCl is saturated with PB?. Some of the binding measurements were carried out in ethanol-water mixtures.  相似文献   

15.
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   

16.
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   

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.
M. Ilavský  K. Dušek 《Polymer》1983,24(8):981-990
The equilibrium mechanical and optical behaviour of networks prepared from poly(oxypropylene) triols (PPT) and 4,4′-diphenylmethane diisocyanate (MDI) at various initial molar ratios of reactive groups, rH = [OH][NCO], in the range 0.6 < rH < 1.75 have been investigated. The reaction proceeded at 80°C and in the presence of an organotin catalyst to a high conversion of minotirty groups (0.96–1.0). A comparison between experimental and theoretical (based on the theory of branching processes) dependences of the equilibrium modulus, Ge, on rH or on the sol content, ws, led to the following conclusions: (1) In the range rH ? 1 the theory adequately describes experimental dependences, while in the range rH < 1 excess crosslinking takes place, obviously due to the formation of allophanate groups. (2) In PPT networks with a higher molecular weight (Mn = 2630), Ge is about twice the theoretical value for the front factor A = 1, which is interpreted as a contribution due to permanent interchain interactions. (3) In networks of lower PPT (Mn = 708), experimental Ge values lie between the theoretical ones calculated for A = 1 (affine deformation crosslinks) and A = 13 (phantom network). (4) The difference between the eperimental and theoretical moduli for A = 13 may be adequately described by using Langley's concept of trapped entanglement contribution (for networks of the longer triol quite satisfactorily, for those of shorter PPT with some systematic deviation) with the same proportionality constant. (5) For correlations between the experiment and theory a generalized plot of the reduced modulus of the weight fraction of the gel proved to be useful.  相似文献   

19.
It was found that the electrochemical reduction of poly (tetrafluoroethylene) with lithium amalgam leading to the mixture of polymeric carbon and lithium fluoride as solid product is followed by a consecutive electrochemical reaction via a polymeric anion radical, [(?Cn)??]m. The value of n was determined as 4.75 and was found to be independent of the reaction conditions. The charge of this radical is during the reaction neutralised with Li+ ions diffusing through the solid phase to form a chemical compound (?CnLi)?m whose CLi bonds have a localized character. Therefore, this compound does not evolve hydrogen in contact with water as the intercalation compounds CnLi, but hydrolyses to a solid hydrocarbon (?CnH)?m.  相似文献   

20.
C.P. Tsonis  M.U. Hasan 《Polymer》1983,24(6):707-712
Arene chromium tricarbonyls were found to function as effective homogeneous catalysts for the polycondensation of benzyl chloride when they are thermally activated. The catalytic activity and induction period depended upon the nature of the arene attached to the metal. Their activity decreased and induction time increased with respect to the nature of the arene in the order: anisole, toluene, p-xylene and benzene. The alkylation products of a model reaction, catalysed by ArCr(CO)3, were only ortho- and para-substituted. The degree of polybenzyl branching, determined by 1H n.m.r. and DP, was found to increase with increasing reaction temperature and M?n. Although there is experimental evidence that in the initiation step the arene ring remains loosely attached to the metal, a mechanism similar to that of polymerization by a Lewis acid catalyst is proposed.  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