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.
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   

2.
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   

3.
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   

4.
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   

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.
7.
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   

8.
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   

9.
Shaul M. Aharoni 《Polymer》1980,21(12):1413-1422
When plotted against concentration V02, the viscosity η0 curve of isotropic solutions of lyotropic nematic mesomorphic polymers increases according to:
η00[[η]V02+(π4)[η]2(V02)2+K2(lnx)2[η]3(V02)3+…]
in which η0 is the solvent viscosity, K2 a numerical constant, x? is the average molecular axial ratio and [η] the intrinsic viscosity as defined by:
[η]=2x2451ln 2x?1.84+3ln 2x?0.61 (+1415)
At the concentration v12 an anisotropic phase appears and the viscosity curve shows a decrease in slope followed by a change in direction at a peak viscosity ηp at vp2. Upon further increase in v02 the system undergoes a phase inversion and finally turns fully anisotropic at vA2. In the biphasic interval the viscosity is described by:
η0mat 1+(5λ+2)2(λ+1)Vinc
where λ = ηincηmat, the ratio of the inclusions viscosity to the matrix viscosity, and vinc the volume fraction of the inclusions in the system. It must be emphasized that the molecular weight of the polymer in both phases changes continuously with concentration, resulting in commensurate changes in ηince, ηmat and their ratio λ. In the anisotropic region the viscosity first decreases moderately and then increases precipitously with v02, according to:
η00[2x290S(5ln 2x?1.8+6ln 2x?0.61)+2]V021-V25°VE
in which S is an order parameter and V25, VE are volumes swept by the orbits of the flowing rodlike macromolecules. The equations give results in good qualitative and fair quantitative agreement with experimental data in the literature.  相似文献   

10.
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   

11.
12.
Dilute solution properties of poly(phenyl thiolmethacrylate) (PTPh) and poly(o-methylphenyl-thiolmethacrylate) (PTMPh) were studied by gel permeation chromatography, light scattering, osmotic pressure and viscosity measurements in different solvents. Relations between intrinsic viscosity [η], z-average root-mean-square end-to-end distance (r?2)12 and molecular weight M?w were established. The unperturbed dimensions were calculated by different methods using M?w and [η]. PTPh (σ = 2.26) and PTMPh (σ = 2.27) have practically the same flexibility, but both are more flexible than poly(phenyl methacrylate) (σ = 2.50). In our opinion, this is due to the change in the local intramolecular interaction and the higher flexibility of the sulphur-containing side group.  相似文献   

13.
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   

14.
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   

15.
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   

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.
18.
E. Straube 《Polymer》1985,26(1):105-108
A polymer chain consisting of Nr segments with a repulsive interaction (binary cluster integral βr) and Na ? Nr segments with a stronger, attractive and pairwise saturable interaction (βa), which is at the averaged θ-point N2rβr + N2aβa = 0 deviations from the predictions of the two parameter theory: α2R ? 1 ~ δzr < 0 and A2δzr > 0 with δzr ~ βr(NaNr)12. It is shown that the deviations from the universal behaviour are due to the existence of an intermediate length scale NaNr.  相似文献   

19.
The mass transfer rate of hydrogen in tetralin and hydrogenated SRC II liquid was studied in a stirred vessel at 606–684 K and 7.0–13.5 MPa. Experiments were carried out using a newly developed in-situ hydrogen probe made of semi-permeable nickel membrane. The effects of stirrer speed, liquid height to vessel diameter ratio, temperature and pressure on mass transfer rate coefficients were investigated. The experimentally determined Kla values were correlated in terms of power input per unit volume of liquid and liquid height to vessel diameter ratio as follows: kLa = 3.43 × 10?4 (PV)0.8 (HDT)?1.9 Furthermore, the liquid-phase mass transfer coefficient, kl, was found to be of the order of 10?5 m s?1 for low agitator speeds.  相似文献   

20.
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   

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