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.
A new model is proposed for the chain conformation and packing of MDI-butandiol hard segments in polyurethane elastomers. The model is based on the structure of methanol-capped MDI (MeMMe), which we have recently determined by single-crystal X-ray methods. Planar zig-zag -CH2CH2-sections connect successive MeMMe units, which have the same conformation as in the monomer structure. Successive monomer units along the chain are related by a centre of symmetry at the central CH2CH2 bond. The chains are linked together in stacks through CO…HN hydrogen bonds which involve half of the urethane groups. The remaining urethanes are similarly hydrogen-bonded to adjacent stacks, and thus the structure is stabilized by hydrogen-bonding in both directions perpendicular to the chain axis. A triclinic unit cell is proposed for the structure with approximate dimensions a = 5.2 A?, b = 4.8 A?, c = 35.0 A?, α = 115°, β = 121° and γ = 85°. The space group is P1 and the cell contains two repeat units of a single chain.  相似文献   

2.
3.
The limiting viscosity number in polystyrene-cyclopentane system has been determined over the temperature range of θu to θl in which θu and θl are the θ or Flory temperature for the upper and lower critical solution temperatures. The temperature coefficient of unperturbed mean square end-to-end distance observed for the polystyrene (Mw=20×104, MwMn<1·06 and Mw=67×104, MwMn<1·10) in cyclopentane is negligibly small. The observed temperature dependence of the polymer chain dimension over the temperature range of θu=19·6° to θl=154·2°C shows a parabolic curve with a maximum in the neighbourhood of 90°C and is qualitatively interpreted by the free volume theory of polymer solution, which gives a new χ1-temperature function.  相似文献   

4.
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   

5.
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   

6.
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   

7.
Structural ordering changes of a non-crystalline polycarbonate (trade name ‘Bistan AW’) were investigated. Three interference halo signals were recorded on the X-ray diffraction patterns at reflection angles equal to θ1 = 3.1° (d = 14.3 A?), θ2 = 8.5° (d = 5.2 A?), and θ3 = 12.7° (d = 3.5 A?). The areas of peaks at θ1 were found to decrease and those at θ2 to increase with increasing the time of annealing the previously melted polycarbonate samples at a temperature of 190°C. It was concluded that morphological changes were taking place in the polycarbonate prior to its thermal crystallization. These changes may be explained by assuming the existence of two morphological forms of structural ordering in non-crystalline polycarbonate macromolecules. On annealing, macromolecules belonging to the cis-form having a lower degree of structural ordering most probably tend to straighten out and assume a zigzag structure, i.e. all-trans conformation. After straightening out of the chains, these macromolecules link up with areas exhibiting a much higher degree of structural ordering and most probably exist in the trans-form.  相似文献   

8.
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   

9.
The relaxational and retardational properties of poly(propylene glycol) liquids, of nominal molecular weights 400 and 4000, are described. The viscoelastic behaviour of each liquid has been determined over a wide temperature range, using high frequency shear wave techniques operating at 30 and 454 MHz. It is found that the complex compliance J1(jω) is described in terms of the viscosity ν, the limiting high frequency compliance J, the retardational compliance Jr and a characteristic retardation time тr by:
J1(jω)=J+1jωη+jr(1+jωτr)β
where β is a parameter of the retardation time distribution.For the lower molecular weight liquid, JrJ = 17.4, β = 0.45 and тrтm increases with decreasing temperature, reaching a limit of 170 near 0°C. This liquid shows no evidence of polymeric behaviour.For the other, JrJ, β = 0.76, тrтm = 15.4 and is constant over the temperature range investigated. The main difference between the two liquids appears as an additional retardational or relaxational process for the higher molecular weight material which occurs in the initial or low frequency part of the relaxation region. This process is characterized by a single time, but with relaxation time 17 and a stiffness 7 times the values calculated for the first Rouse mode of polymer chain motion.  相似文献   

10.
A combination of steady-state and fluorescence decay techniques permits one to measure the dynamics of end-to-end cyclization of a polymer chain substituted at both ends with pyrene groups. In the limit of low concentration, the rate constant for cyclization, kcy, can be identified with the slowest relaxation rate τ1?1 of a Rouse—Zimm chain. Experiments are reported which allow kcy to be examined for two chain lengths of polystyrene substituted on both ends with pyrene groups. These chains have M?n = 9200 and 25 000 (M?wM?n ? 1.15). Added unlabelled polystyrene polymer [PS] causes k?cy to decrease in cyclohexane just above the θ-temperature, whereas in toluene, a good solvent, kcy is largely unaffected, even at [PS] concentrations of 50 wt%. These results are explained in terms of frictional effects—hydrodynamic screening—dominating in the poor solvent, whereas other factors tend to have offsetting effects in the good solvent.  相似文献   

11.
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   

12.
C.L. Choy  F.C. Chen  E.L. Ong 《Polymer》1979,20(10):1191-1198
The linear thermal expansitivity of oriented high-density polyethylene (HDPE) and polypropylene (PP) with draw ratio between 1 and 18 has been measured between 120 and 300 K. The expansivity perpendicular to the draw direction z?) increases with λ, because of the alignment of the crystallite chain axes along ?; the expansivity parallel to z?) decreases very sharply with λ, becoming negative at about λ = 3 for HDPE and λ = 7 for PP — a consequence of the negative expansivity of the crystalline phase along its chain axis and the constraining effect of the stiff intercrystalline bridges. A model treating the drawn crystalline polymer as a composite consisting of partly aligned crystallites embedded in an isotropic amorphous matrix is adequate for explaining the behaviour of α, whereas a parallel-series model can give reasonable estimates of α for ultra-oriented samples, which is largely independent of λ and decreases with increasing temperature. However, neither model can account for the behaviour of α at low draw ratio.  相似文献   

13.
J.C. Radon  L.E. Culver 《Polymer》1975,16(7):539-544
The effects of frequency and temperature on fatigue crack propagation rate in poly(methyl methacrylate) and polycarbonate have been studied using centrally notched plate specimens cycled in tension between constant stress intensity limits. Crack growth was monitored at frequencies between 0.1 Hz and 100 Hz and at temperatures between ?60°C and 40°C. A linear relationship between the cyclic crack growth rate d(2a)dN and appropriate levels of toughness, K, has been proposed: d(2a)dN = A?α, where ? = (λ ? λth)(K21C ? K2max), λ = K2max ? K2min, λth is the threshold limit and A and α are constants. Also, the influence of mean stress intensity was briefly discussed.  相似文献   

14.
A. Campos  B. Celda  J. Mora  J.E. Figueruelo 《Polymer》1984,25(10):1479-1485
Intrinsic viscosities, [η], second virial coefficients, A2, preferential solvation coefficients, λ, and binary interaction potential as measured by light scattering, g12, for the system n-undecane(1)/butanone(2)/poly(dimethylsiloxane) (3) have been determined at 20.0°C. The system shows cosolvent character, as the inversion in λ and the maxima in A2 and in [η], at ø10?0.65, seem to indicate. (g13sg23) and the ternary interaction potential, gT, and its derivatives on system composition, (?gT?u1)ø3→0 and ?gT3)u13→0, have been evaluated. Global interaction parameters, χm3, have also been evaluated and a critical analysis on the approximations usually followed for χm3 calculations is undertaken.  相似文献   

15.
Kock-Yee Law 《Polymer》1984,25(3):399-402
The effect of solvent vapour on the properties of vapour-swollen vinyl chloride-vinyl acetate (8317) copolymer has been studied by a fluorescence probe, a p-N,N-dialkylaminobenzylidenemalononitrile derivative (1). Results show that the fluorescence quantum yield (Фf) of 1 in vinyl chloride-vinyl acetate (8317) matrix decreases by a factor of ≈ 10, an indication of the increase in free volume or in polymer chain mobility, upon vapour swelling. The variations of Фf observed in various swollen matrices, which correlate only with the density of the swelling solvent, indicate that there is a profound vapour effect on the properties of swollen polymer. A density effect on the mobility of polymer chains in swollen polymer is proposed.  相似文献   

16.
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   

17.
To model the reversible novolac polymerization, five reactive species A to E have been defined. Molecules having bound CH2OH (Qn) are distinguished from those without it (Pn) and it is assumed that molecules of Qn do not have more than one bound CH2OH group. A kinetic model has been written and, based upon it, balance equations for molecules of novolac polymer in batch reactors have been derived. Based upon our earlier studies, the phenomenon of molecular shielding has been neglected. As a result, the reactivities of the ortho and para positions of phenol which are available in the literature could be used. The kinetic model for the molecular weight distribution (MWD) of reversible novolac polymer formation thus involves only one parameter. The study of the MWD of novolac polymer reveals two very important design variables: the phenol-formaldehyde ratio, [P]0[F]0, in the feed and the vacuum applied on the reactor. As the [P]0[F]0 ratio is increased, the breadth of the distribution is found to increase and it undergoes a maximum at [P]0[F]0 ? 1.4 for the set of rate constants chosen. At this ratio, the chain length average molecular weight is also found to be the largest. Industrially, the [P]0[F]0 ratio used in producing novolac polymer is 1.67 and it is usually desired that the polymer be linear with minimal branching. On application of vacuum, for a given time of polymerization, the chain length molecular weight is found to increase when the results are compared with those of batch reactors. The breadth of the distribution is also found to reduce thus giving a lower polydispersity index of the polymer formed.  相似文献   

18.
The determination of the crystalline structure of oriented fibres of poly(trimethylene terephthalate) is described. The unit cell is triclinic with the following parameters: a = 4.6A?, b = 6.2A?, c = 18.3A?, α = 98°, β = 90°, γ = 112°. Each cell contains two monomers of one polymer chain. Both methylene bonds are in the gauche conformation. The chain conformation and the packing of chains within the unit cell are discussed in detail and compared with other chemically similar materials, both monomeric and polymeric.  相似文献   

19.
Samples of poly(ethylene terephthalate) (PET) modified with small amounts of trimesic acid groups and hence containing long chain branching have been prepared. From the content of trifunctional modifier and from the experimental value of the extent of reaction, the weight-average molecular weight M?w and branching density B?w have been calculated, assuming that all the end-groups are equally reactive and intramolecular reactions are absent. The values of M?w and B?w have been correlated with the experimental values of intrinsic viscosity [η] and the Newtonian melt viscosity η0. General relations of the following type have been obtained:
f1([η], Mw, Bw) = 0; f20, Mw, Bw) =0; f30, [η], Bw) = 0; f40, [η], Mw) = 0;
In particular, [η] and η0 increase on increasing M?w and decrease on increasing B?w, but, at equal [η] values, η0 increases with B?w. Through the last relation, the reliability limits of which should be experimentally checked, and from measurements of [η] and η0, it is possible to calculate M?w of a branched PET.  相似文献   

20.
Hydrocracking of benzyl phenyl ether and diphenyl ether has been carried out over a series of iron catalysts (Fe2O3, Fe2O3Al2O3, Fe2O3ZnO, Fe2O3ZrO2, Fe2O3MgO and Fe2O3SiO2) and reference catalysts (CoOMoO3Al2O3, NiOMoO3Al2O3 and SiO _Al2O3) to search for active and selective catalysts, and to elucidate the catalyst properties of relevance for CO bond cleavage. Among the iron catalysts, Fe2O3ZnO, Fe2O3ZrO2 and Fe2O3MgO exhibited relatively high activities and selectivities. The important property of catalysts relevant to hydrocracking of the ethers is the ability to accomplish hydrogenation. Acidic properties of the catalysts cause condensation and rearrangement of the reactants. The features of the iron catalysts are compared with those of molybdenum catalysts, and the reaction mechanisms discussed.  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