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.
Two coking coals, a caking and a non-caking coal are examined in a Bruker pulsed 1H n.m.r. spectrometer in the temperature range 293–730 K. One coking and the caking coal are oxidized in air at 383 K for 13 days. Temperatures of signal appearance and loss are noted as well as the temperatures of minimum signal half-peak width (ΔH12). There occurs no change in the above three temperatures with oxidation of the coals. The variation of (ΔH12) with temperature of the coal is also measured. Changes in (ΔH12) are more pronounced for the caking coal. The softening and solidification temperatures are below and above, respectively, those reported using the Gieseler method. Values of (ΔH12) increase beyond the minimum value as the fluidity continues to increase. This may be caused by an increase in average molecular weight of constituent molecules and/or an increasing concentration of free radicals in the fluid phase. This experimental approach may afford a new method to characterize coals which are to be used in liquefaction processes.  相似文献   

2.
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   

3.
C.J. Farrell  A. Keller  M.J. Miles  D.P. Pope 《Polymer》1980,21(11):1292-1294
The nature of chain extension of high-molecular weight atactic polystyrene in an elongational flow field was investigated. In particular, we studied the sudden onset of high extension which, according to theory, is governed by the expression ??cτ = 1. The molecular weight dependence of ??c (the critical strain rate at which high extension occurs) was examined, and the corresponding relaxation time, τ, which would be relevant to the above expression was considered. The experimental measurements of absolute values and the functional dependence on molecular weight of the relaxation times demonstrates that it is the relaxation time associated with the Zimm non-free-draining model which is appropriate to this coil-stretch transition.  相似文献   

4.
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   

5.
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   

6.
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   

7.
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   

8.
A synthetic sequence is described for the preparation of polystyrenes in the molecular weight range (M?n) 3 × 103 to 2 × 104 having terminal azodicarboxylate functionality with one functional group per polymer chain. The polymer end groups are characterized spectroscopically at each step in the synthetic sequence. The concentration of azodicarboxylate groups on the polymers is determined spectro-scopically and compared with the M?n of the polymers as calculated from initiator/monomer ratio or as measured by g.p.c. analysis.  相似文献   

9.
10.
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   

11.
A.I. Hopwood  H.J. Coles 《Polymer》1985,26(9):1312-1318
Magnetic and electric fields have been used to determine the splay (k11) and bend (k33) elastic constants in a series of polymer/monomer liquid crystal solutions using the well known Freedericksz transition technique. Measurements have been carried out as a function of concentration and temperature. The polymer used was a smectogenic polysiloxane side chain liquid crystal with both cyanobiphenyl and benzoate ester side groups. The monomeric solvent was the nematogen 4-n-pentyl-4′-cyanobiphenyl. All of the solutions studied were nematogenic up to a concentration of 40% w/w. It has been shown that k11, k33 and k33k11 all decrease with increasing polymer concentration and that at high enough concentrations k33k11 tends to become independent of temperature. The implications of these results are discussed in terms of the performance of the most common liquid crystal display, i.e. the twisted nematic device.  相似文献   

12.
A flow microcalorimeter designed to measure the heat of mixing of dilute polymer solutions is described. The instrument is sensitive to steady state heating rates of ~10 μJ/sec. Measurements of heats of mixing of solutions of differing concentrations of n-hexane and cyclohexane are reported and are compared with recommended data of McGlashan and Stoeckli. Values of:
K1=limV2→ 0
(H?1 ? H?01RTv22 are obtained for four polymer—solvent systems: polyisobutylene—benzene, 0.22; polystyrene (PS)—cyclohexane, 0.33; PS—n-butyl acetate, ?0.06 all at 25°C; and PS—toluene, ?0.05 at 40°C. Various theoretical calculations of second virial coefficients A2 made with use of the calorimetric data are compared with previously measured A2 for the first two mixtures.  相似文献   

13.
High resolution neutron scattering experiments have been used to observe the diffusive motion of low molecular weight linear and cyclic poly(dimethyl siloxane) molecules in dilute solution in deuterated benzene. Diffusion coefficients (D) and hydrodynamic radii (RH) have been compared with values obtained by light scattering for higher molecular weight samples and with radii of gyration (Rg) obtained by small-angle neutron scattering. While the ratio DringDchain is close to the predicted value of 0.85, the ratio RgRH falls below the theoretical value for both ring and chain molecules. The scattering curves show effects arising from both centre of mass diffusion and internal molecular motion, and the observed inverse correlation times are compared with calculated behaviour as a function of scattering vector, Q.  相似文献   

14.
K. Dodgson  D. Sympson  J.A. Semlyen 《Polymer》1978,19(11):1285-1289
A preparative gel permeation chromatographic (g.p.c.) instrument has been constructed and used to separate broad fractions of cyclic poly(dimethyl siloxanes) into sharp fractions with heterogeneity indices M?wM?n = 1.05 ± 0.02. The number-average molecular weights M?n of the cyclic polymer fractions obtained were as high as 50 000, corresponding to number-average numbers of skeletal bonds n?n up to 1300. The concentrations of linear poly(dimethyl siloxanes) in all but the highest molecular weight cyclic polymer fractions prepared are believed to be negligible. The preparative g.p.c. instrument was also used to obtain some sharp fractions of linear poly(dimethyl siloxanes).  相似文献   

15.
A Monte Carlo method has been devised for calculating the conformation-dependent properties of cyclic poly(dimethyl siloxanes) (PDMS), using Flory, Crescenzi and Mark's rotational isomeric state model. Calculated values of the mean-square radii of gyration 〈s2r〉 of ring molecules unperturbed by excluded volume effects and containing 8–100 skeletal atoms are compared with the 〈s2l〉 values for the corresponding unperturbed chain molecules. Exact enumeration methods were also employed for rings [(CH3)2SiO]w2 with w ? 24 and the results found to be in close agreement with those obtained by the Monte Carlo method. The ratio 〈s2l〈s2r was found to attain limiting values close to 2.0 for w > 30, in agreement with theoretical predictions.  相似文献   

16.
The tacticity and number average sequence length of like (n?o), meso (n?m) and racemic (n?r) acrylonitrile (AN) triad units in polyacrylonitrile (PAN) prepared in both water and water—acetone (2:1 v/v) media and AN-3-chloro, 2-hydroxypropyl acrylate/methacrylate, AN-2-bromoethyl methacrylate and AN-2-chloroethyl acrylate copolymers have been calculated using 13C n.m.r. spectra of the polymer solutions concerned at a field strength of 24.99 MHz. The spectra reveal that PAN prepared in water medium has a greater percentage (33.4%) of isotactic units than PAN prepared in water-acetone (2:1 v/v) medium (28.3%). The tacticity distribution of AN sequences in PAN and the copolymers is found to be random (n?m?n?r?2.0) and the number average sequence length of AN sequences in a copolymer containing 14.8 mole% of 3-chloro, 2-hydroxypropyl methacrylate was 15.2.  相似文献   

17.
J. Maxfield  I.W. Shepherd 《Polymer》1975,16(7):505-509
The Raman spectrum of poly(ethylene oxide) (PEO) M?w = 3 × 106 and 6 × 103 has been measured in bulk as a function of temperature and in aqueous and chloroform solution as a function of solvent concentration. The spectral features are assigned to particular isomeric configurations and the changes on melting are found to be consistent with a helix-coil transition. In both solvents the ordered nature of the polymer is largely lost at a critical concentration, approximately 50% by weight; residual ordering in dilute aqueous solution is lost and the random coil configuration attained only above the melting temperature. The wavenumber change of the 862 cm?1 mode as a function of water concentration showed formation of a complex involving three water molecules and was consistent with a simple H-bonding model. The differences between the spectra in the two solvents are explained by this H-bonding. The results are in general agreement with n.m.r. work.  相似文献   

18.
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   

19.
M.J. Miles  A. Keller 《Polymer》1980,21(11):1295-1298
Using a planar-flow cross-slot device the retraction of the molecules from full extension was followed through birefringence observations and the relaxation time associated with this process was found to be considerably larger than that for the extension from the coiled state found in Part 1. Thus in the first approximation a hysteresis is demonstrated together with further details (possibly a double relaxation) of the contraction process. Use of this flow geometry also facilitated exploitation of the relation between Mw and ??c of Part 1 for the method of distinguishing molecular weights in a mixture.  相似文献   

20.
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   

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