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 46 毫秒
1.
In this paper, we present dielectric results for samples of POE (polyethyleneoxyde) divided into three classes. Compounds of low molecular weight (Mw < 1000) show very weak absorption at lower temperatures. At higher temperatures (below the melting point), there appears a very important peak (stronger for lower Mw) which corresponds to the supercooled phase. Compounds of intermediate molecular weight (1000 ≤ Mw ≤ 105) show a β relaxation near 250 K (bisMw ? 3000) which is due to the movement of chain-ends in the amorphous phase. This process increases with the importance of that part of the structure formed of shorter lamellae, thus explaining the marked diminution of this absorption for higher molecular weight compounds (the shorter lamellae cannot be obtained for Mw > 4000). The α relaxation (obtained for T > 300 K) is perturbed by an important conduction; however, we have found a peak and a shoulder at lower frequencies, due to shorter and longer lamellae respectively. The shoulder is also present for high molecular weight compounds (Mw ≥ 0.1–106). At lower temperatures, they display a very broad γ absorption, superposed at higher frequencies with a β relaxation. The nature of this β relaxation is different from that observed for the intermediate molecular weight compounds.  相似文献   

2.
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   

3.
Molecular weight distributions of the cationic polyelectrolyte poly(2-trimethylammonium ethyl methacrylate chloride) were determined via size exclusion chromatography (SEC) for a very broad molecular weight range (7.7 × 103 ? Mw ? 1.1 × 107). Non-size exclusion effects (ion inclusion, ion exclusion, adsorption) were controlled by special surface treatment of the stationary phase (DMAE-FractosilR, E. Merck) and by proper selection of the mobile phase. A non-linear effective molecular weight calibration procedure was applied to account for the high polydispersity of the polyelectrolyte standard polymers. Mw data from SEC experiments agree with results from light scattering, and the molecular weight distributions obtained from SEC and sedimentation velocity analysis compare well.  相似文献   

4.
Linear and branched bisphenol A polycarbonate (PC) samples were characterized by their average molecular weights, Mn and Mw, polydispersity degree q = Mw/Mn, and branching degree gv. The weight fraction of microgel was also determined for branched samples. The samples were amorphized and densities were measured at 23°C to obtain the values of specific volume, vsp. The dependence of vsp on molecular characteristics is described by the multivariable power function Δvsp = AspMxaqapx gvab, where Δvsp = vsp ? vsp,∞, and Asp, a, apx and ab are constants. It has been confirmed that a = ?1, apn = 0 and apw = 1. It has also been found that the branching exponent ab significantly depends on microgel content. The relationships found for PC should, in principle, be valid for other polymers. Examples based on literature data are given for linear polyethylene and polydimethylsiloxane.  相似文献   

5.
The scattering function P0 for classical elastic light scattering is specified for several molecular weight distribution functions (Schulz-, Square root-, Maxwell-, Normal-, Poisson-, Logarithmic-normal-distribution function). Precision measurements on anionic polymerized polystyrene from Pressure Chemical Co. (Mw = 2 · 106 and Mw = 6,7 · 106) and radical initiated polystyrene were performed in transdecalin with the light scattering photometer Fica 50. Comparison of experimental results with theoretical curves indicate that the anionic polystyrenes exhibit a broader distribution than given by Pressure Chemical Co. The distribution of the radical polystyrene conforms with distributions found by other methods.  相似文献   

6.
An investigation has been carried out into the effect of the fractional composition on the rheological (flow and elastic) properties of a system, using mixtures of polybutadienes with narrow molecularweight distribution (MWD). For mixtures of high-molecular-weight components, the initial Newtonian viscosity is determined by the weight-average molecular weight: η0Mαw; when low-molecular weight components are introduced, it is also determined by the MWD moment ratio. The characteristic relaxation time of a system is determined by the z-average molecular weight: θ0Mα1z, and in the general case α1α. A new model has been proposed to explain the non-Newtonian phenomenon as a consequence of the existence of a molecular-weight distribution. According to this model, as the shear rate increases the high-molecular-weight components gradually (at their critical rates) pass over to the high-elastic state. Therefore, at high shear rates, their contribution to viscous losses of a polydisperse polymer is associated with their behaviour as a viscoelastic filler in a viscous liquid.  相似文献   

7.
Flash pyrolysis of radical and anionic polyacrylonitriles between 400 and 800° and in the absence of oxygen leads to various degradation products, the most characteristic being cyanhydric acid,. cyanogen, acrylonitrile and two dimers α-methyleneglutaronitrile and a cyanopicoline. Two important parameters have been studied: molecular weight (4 × 103 < Mw < 1·6 × 106) and branching density (0 < N/X < 0·6). At 400°, the yields in acrylonitrile and α-methylene glutaronitrile increase with branching density. All the experimental results may be interpreted on the basis of random chain scission, hydrogen transfer reactions and cyclization of the nitrile groups.  相似文献   

8.
A procedure is described for the determination of the molecular weight distribution (MWD) by dynamic light scattering, in which no restrictions with respect to the scattering system are necessary. The procedure involves the construction of one single curve from a system of experimental curves. This curve is suitable for the determination of the MWD since all effects are eliminated which are in general not exactly describable by theory (such as concentration effects, intramolecular motion, scattering function) and which therefore can systematically distort the results for the MWD. The procedure is illustrated and tested for a polystyrene with Mw ? 18·106 . in toluene at 20°C.  相似文献   

9.
A method for evaluation of the type of average, which is experimentally obtained for a given property of polydisperse polymer, is described. A multivariable power function
P=APMXaqapx
where P is the polymer property, Mx is the x-average molecular weight, q is the polydispersity degree, Ap, a and apx are constants, and the apx = 0 criterion (apx being the polydispersity exponent) is used for this purpose.  相似文献   

10.
The photoxidative degradation of PVC was studied by i.r. and u.v. spectroscopy and gel permeation chromatography. It was found that the photoxidative degradation of unprocessed PVC is an auto-accelerating chain scission process. Carboxylic groups were found to be the main product formed during degradation. The molecular weights Mn and Mw both decreased, but the molecular weight distribution widened with increasing length of exposure. Single additives, such as calcium stearate. Wax E and a solid organotin stabilizer, altered the rate but not the mode of the degradation. Combination of the three additives changed the mode of the photoxidation from auto-accelerating to constant rate of degradation. Processing at 170° with the combined additives increased the rate constants.  相似文献   

11.
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   

12.
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   

13.
The kinetics of the thermal polymerization of styrene have been studied over the range 60–250°. The overall energy of activation was 86 ± 2 kJ/mole, a value identical to that obtained for the thermal polymerization of styrene in diethyl adipate. As expected, the molecular weight of the polymer decreases with increase in the temperature of the polymerization, and the ratio MwMn becomes greater than 2 for polymer formed at above 140°. The plot of log (1Mn) against (absolute temperature)?1 can be represented by two straight lines yielding 24.5 and 32,0 kJ/mole for the activation energies at temperatures below 120° and above 140°, respectively. The former value is in keeping with the molecular weight being controlled by chain transfer with monomer; the latter value would be that expected if the termination process controls the molecular weight of the polymer. Mark Houwink relationships between intrinsic viscosity and Mn and Mw have been found to apply to polymer samples when the molecular weight averages were determined by osmometry and by light scattering. However, deviations were found for low molecular weight material when measured using gel permeation chromatography. The K values were considerably lower, and the α values higher than reported in the literature.  相似文献   

14.
Gel permeation chromatography (GPC) and viscometry (V) methods have been combined for determination of long-chain branching in bisphenol-A polycarbonate (PC) by means of a branching factor gv = Mvg1/Mv1, where Mvg1 and Mv1 are the apparent viscosity-average molecular weights calculated from GPC data and from intrinsic viscosities [η] of the samples respectively. A linear dependence of gv on molar % of branching agent has been obtained. The GPC data on PC samples have also been applied for verification of an earlier [η]?M relationship for branched polydisperse polymers.  相似文献   

15.
The polymerization of acrylamide, initiated with permanganate/oxalic acid, has been studied at 35 ± 0.2 in an aqueous medium under nitrogen. Samples of polyacrylamide were fractionated by the triangular fractionation method using methanol as non-solvent. Molecular weights of the fractions have been determined by viscometry and osmometry. Integral and differential distribution curves were plotted using the fractionation data. The narrow molecular weight distribution for high conversion polymers has been discussed. For fractionated samples of polyacrylamide in water at 30°, the equation [η]ml/g = 6.5 × 10?3Mn0.82 is applicable for the molecular weight range 4 × 104 to 127 × 104. This equation is very similar to the equation [η]ml/g = 6.31 × 10?3M/0.80 of Scholtan. Other parameters (osmotic second virial coefficient and unperturbed dimension) have also been evaluated.  相似文献   

16.
We have tried to find the pair of z, h(z) functions most conveniently accounting for the variation of A2 with the molecular wieght, M. The selected z, h(z) pair led to a relation of the type A2M12 = f(M12) allowing determination of the unperturbed dimensions (L02/M)12 and the B coefficient characterizing the polymer-solvent interactions. The results obtained with various polymer-solvent pairs are compared to those arising from the Stockmayer-Fixman modified equation.  相似文献   

17.
18.
Weight-average molecular weights for PDMS fractions were determined by light scattering and the Archibald method in the range of 3000–300,000; the [ν]-M? relationship for toluene, 25°, was found to be
[?]=1.87 × 10?2Mw03658,cm3/g
This relationship was compared with those previously used. No upward curvature of the linear relation, log [ν] = ?(M), at lower molecular weights was observed. Over the whole range of molecular weights, the [ν]-values in toluene were higher than [ν]θ-values in bromocyclohexane at 28°.  相似文献   

19.
Photon-correlation spectroscopy has been used to measure the diffusion coefficient D and first-mode intramolecular relaxation time τ1 of polystyrene in a theta solvent, cyclohexane at 34.5°C. Measurements were made on five narrow fractions (Mw = 2.88 × 106 to 9.35 × 106) as a function of concentration c, in the dilute regime. D varied linearly with c, D = Do (I + kDc), with Do = (1.4 ± 0.2) × 10?4Mw?(0.508±0.007) cm2 s?1. Although the values obtained for τ1 were reproducible to within 5%, no systematic variation with c was detected. The results are fitted by the relation τ1 = (7.7 ? 0.3) × 10?8Mw(1.42±0.05) μs, which agrees well with the theoretical expression of Zimm for the non-draining bead-and-spring model, modified for the light-scattering case.  相似文献   

20.
The phase relationships of poly(N-vinyl-3,6-dibromo carbazole) (PVK-3, 6-Br2) were examined for four solvents, viz, o-chlorophenol, p-chloro-m-cresol, o-dichlorobenzene and bromobenzene. Upper critical solution temperatures (UCST) have been determined for solutions of PVK-3,6-Br, fractions in o-chlorophenol and p-chloro-m-cresol over the molecular weight range Mw × 10?4 = 125.0 to 4.8. The Flory temperature, θ, obtained from UCST for the PVK-3,6-Br2/o-chlorophenol and PVK-3,6-Br2/p-chloro-m-cresol systems are 66.0 and 112.9°C, respectively. The θ-temperatures were checked against molecular weight and viscosity data to determine the Mark-Houwink equations for these two theta solvents, with satisfactory agreement. The relations are
[ν] = 27.54 × 10?10 × M0.50w (o-chlorophenol, 60.0°C
[ν] = 30.24 × 10?10 × M0.50w (p-chloro-mcresol, 112.9°C
The characteristic ratio C = 〈R20/nl2 was found to be 16.6 in o-chlorophenol at 60.0°C and 17.6 in p-chloro-m-cresol at 112.9°C. The value of the characteristic ratio C of PVK-3,6-Br2 is of the same order of that for poly(N-vinyl carbazole). This indicates that the bromine atoms at the 3 and 6 (meta) positions have only an inappreciable effect on the hindering potential for rotation about the CC bond. This agreement of C for both polymers may also be taken as indicating that the effect of interaction between polar groups at the m-position on the hindering potential for rotation is small. The phase diagrams of PVK-3,6-Br2 obtained in o-dichlorobenzene and bromobenzene seem to be characteristic of organized phase structures such as those found in systems exhibiting thermoreversible gelation. Light scattering measurement on PVK-3,6-Br2 dissolved in o-dichlorobenzene, a gelation promoting solvent, and tetrahydrofuran, a very good solvent, strongly indicate that the macromolecular species in o-dichlorobenzene contain some extent supermolecular structures (aggregates, association of chain segments, etc.). These characteristic structures of PVK-3,6-Br2 in o-dichlorobenzene and bromobenzene at 25°C are also characterized by high values of the Huggins' constant k′; for tetrahydrofuran solutions, the k′ values were in the range normally found for many good solvent-polymer systems.  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