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.
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   

2.
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   

3.
The methods of isothermal and dynamic thermogravimetry have been used to study the degradation of poly(N-vinylcarbazole) (PNVC). The multiple heating rate method has been used, as a dynamic method, to obtain kinetic parameters. A linear relationship between the activation energy. E, and Mw?1 (Mw being weight average molecular weight) was found. From isothermal experiments, a temperature was found for which E was independent of molecular weight. We could then refer to a degradation characteristic- temperature of the polymer. On the other hand, altering the heating rate leads to changes in the values of E for each molecular weight indicating two kinds of scission: one occurs in the backbone, producing mainly monomer; in the other, both side-group and backbone scissions occur producing different products.  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.
We report here a method which affords the magnitude of the characteristic parameters of a macromolecular chain starting from a polydisperse sample. A statistical treatment of viscosity measurements made on the small fractions of a GPC elution wave enables us to assign to each of them its Mn value. By numerical calculations it is then possible on the one hand to evaluate the Mark-Houwink constants, MnandMw, the f(M) function and on the other hand to estimate according to a hydrodynamic model, the unperturbed geometrical dimensions. The method is tested for polystyrene samples under conditions which allow the dissolution of many polymers, viz in tetrachloroethane (TCE) at 50° and N-methylpyrrolidone at 85°.  相似文献   

6.
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   

7.
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   

8.
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   

9.
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   

10.
The viscous flow parameters for dilute PVC tetrahydrofuran solutions were determined in the temperature range 10–45°. The temperature dependence of the viscosity of the solutions obeys an Arrhenius-type equation. The concentration and the molecular weight dependence of the pre exponential factor of this equation is given by the following relationship:
A=A0·exp[minus;Ka(c·Mw)2]
.Both the values of the viscous flow parameters and the mathematical expression of the pre-exponential factor indicate, by comparison with other results, some stiffness of PVC chains in tetrahydrofuran.  相似文献   

11.
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   

12.
A series of titanates which have perovskite-like arrangements and are isostructural with [CaCu3](Mn4)O12 have been synthesized. The total charge of the A sites can be modified (1) by substituting the Ca2+ cations with monovalent ones and the tetravalent manganese cations of the B sites by a mixture of (Ti4+ + M5+) in which M = Ta, Nb, Sb, or (2) by substituting the Ca2+ cations by a combination of cations plus vacancies. In this case, if the total charge of the A sites is 2, one obtains compounds such as [Th4+1212Cu3](Ti4)O12 and [T3+2313Cu3](Ti4O12 (T = rare earth); on the contrary, if the charge is less than 2, then one has to compensate it by changing that of the B sites. This leads to compounds such as [□Cu3](Ti2M2)O12 (M = Ta, Nb, Sb).  相似文献   

13.
The theory of the potentials of the cells with CaF2-type electrolyte and electrodes based on oxides of alkaline earth metals
O2Pt|〈M1O〉|CaF2|〈M11O〉|Pt, O2
is proposed. The limits of utilization of CaF2 as a function of partial oxygen pressure over electrodes are considered. The conditions of reversibility of the cells are analyzed for obtaining the thermodynamic parameters of the cell reactions under study. The theoretical analysis and experimental results demonstrate the possibilities of using such cells for thermodynamic study of new classes of refractory compounds based on oxides of alkaline earth and rare-earth or transition metals.  相似文献   

14.
Several linear polymeric ferrocene compounds with Mn in the range 1200–3000 have been compounded with a cable-grade, DIOP-plasticized PVC at concentrations of 1–5%; two non-polymeric ketoferrocenes and a heteroferrocene have been included for comparison. Moulded sheets, prepared from the compounded materials, have been evaluated for compatibility by visual inspection, from analyses prior to and after a heat aging treatment, and from stress-strain data. Best performance with respect to all three compatibility criteria has been observed for a poly(ferrocenylenemethylene) with Mn = 2400, which shows excellent compatibility up to a loading of 9%. Good performance has also been demonstrated for a number of related polymers substituted at the bridging carbon atom. These findings suggest promising applications for selected polymeric ferrocenes as multifunctional additives for PVC-based engineering and biomedical materials.  相似文献   

15.
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   

16.
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   

17.
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   

18.
We propose the following empirical relationship between the intrinsic viscosity of a polymer and its molecular weight M.
{[η]?[η]θ/[η][η]θ=?Δρ+A′M12
[η] and [η]0 are the intrinsic viscosities in a good solvent and in θ conditions respectively. Δ?, and A′ are constants characteristic of a system polymer-solvent. This relationship is valid for PS and PMMA in various good solvents and for a range of molecular weight from 3000 to 250,000. It is in this range that the Mark-Houwink-Sakurada equation is least applicable.  相似文献   

19.
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   

20.
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