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 10 毫秒
1.
The homopolymerizations of methyl methacrylate (MMA) in N-methyl-2-pyrrolidone (NMP) and of N-vinyl-2-pyrrolidone (NVP) in methyl isobutyrate (MiB) were investigated at 60°C using AIBN as initiator. Reaction rate and dynamic viscosity of the medium were determined as function of monomer dilution. It follows from the estimated monomer exponent n that MMA reacts in NMP faster (n = 0,739) and NVP in MiB slower (n = 1,495) in comparison with reaction rates obtained from the direct proportionality of monomer concentration. The observed deviation was quantitatively separated in partial deviations caused by chain initiation, propagation and termination. It was found that the viscosity effect on chain termination plays a dominant role in the global solvent effect. Chain propagation is influenced by electron donor-acceptor interactions between monomers and solvent.  相似文献   

2.
Initiation constant 2kd · f, dissociation constant kd, and efficiency factor f of AIBN were measured at 60°C for two model homopolymerization systems: methyl methacrylate (MMA) in N-methyl-2-pyrrolidone (NMP) and N-vinyl-2-pyrrolidone (NVP) in methyl isobutyrate (MiB) as the solvent. The stable Banfield radical was used as inhibitor. A pronounced solvent effect was observed on the side reaction of the inhibitor with the monomers. Particular solvent sensitivity of monomers in this reaction was identified, whereby opposite solvent properties of NMP and MiB as well as of MMA and NVP were found. The initiation constant 2 kd · f depends mainly on the dissociation constant of the initiator kd in regard to its solvent dependence. The efficiency factor f is influenced by monomer concentration, stability (reactivity) of primary radicals and by viscosity of the reaction medium (solvent cage effect).  相似文献   

3.
Pulsed-laser polymerization (PLP) in conjugation with molecular weight distribution (MWD) measurement has emerged as the method of choice for determining the propagation rate coefficient kp in free-radical polymerizations. Detailed guidelines for using this technique (including essential internal consistency checks) and reporting the results therefrom are given by the authors, members of the IUPAC Working Party on Modeling of kinetics and processes of polymerization. The results for PLP-MWD kp measurements from many laboratories for bulk free-radical polymerization of styrene at low conversions and ambient pressure are collated, and are in excellent agreement. They are therefore recommended as constituting a benchmark data set, one that is best fitted by (the confidence ellipsoid for the Arrhenius parameters is also given). These benchmark data are also used to evaluate the merits of several other methods for determining kp; it is found that appropriately calibrated electron paramagnetic resonance spectroscopy appears to yield reliable values of kp for styrene.  相似文献   

4.
The radical copolymerization of cyclododecyl acrylate (CDA) with styrene (St) or acrylonitrile (AN) was studied in bulk, benzene, tetrahydrofuran, or dioxane at 60°C, and was compared with that of cyclohexyl acrylate (CHA). In the copolymerization with St, the change in the ester groups or the used solvents influenced the reactivity of both acrylates in the same manner as their homopolymerization rates. Approximately, CDA and CHA behaved like methyl methacrylate in the copolymerization with St. Contrary, in the cases of copolymerization of CDA with AN, peculiar results were obtained and interpreted in terms of characteristic association states of the monomers in the solutions. In benzene solution both the monomer reactivity ratios, r1 and r2, are larger than unity (M1 = AN, M2 = CDA; r1 = 1,7, r2 = 2,0).  相似文献   

5.
Binary systems of methyl methacrylate (MMA)/N-vinyl-2-pyrrolidone (NVP) and MMA/N-methyl-2-pyrrolidone (NMP) with NMP as saturated model of NVP and of NVP/methyl isobutyrate (MiB) with MiB as saturated model of MMA were investigated by means of IR and NMR spectroscopy. Investigations were carried out at room temperature or at 60°C in CHCl3 (IR) or CDCl3 and C6H12/C6D12 (NMR). It can be concluded from IR and NMR spectra that the polarity of MMA increases in the presence of NVP and the polarity of NVP decreases in the presence of MMA. Equilibrium constants K of complex formation were determined to: K = 0,169 ± 0,037 L · mol?1 for MMA/NVP at 30°C, 0,112 ± 0,024 L · mol?1 for NVP/MiB at 60°C and 0,125 ± 0,030 L · mol?1 for MMA/NMP at 60°C.  相似文献   

6.
The radical homopolymerization of cyclododecyl acrylate at 60°C in benzene, tetrahydrofuran, or dioxane was studied and compared with that of cyclohexyl acrylate. The difference in the polymerization rates of the acrylates could not be interpreted in terms of the polar and steric substituent constants of ester groups in Taft's equation. Remarkable solvent effects on the polymerization rates for both acrylates were also found in the same manner and ascribed to the polar nature of the used solvents.  相似文献   

7.
Free-radical copolymerization of butyl acrylate with methyl methacrylate was carried out at 50°C in a 3 mol/L benzonitrile solution. Differences between the apparent reactivity ratios determined in this work with those previously reported in bulk, toluene and benzene indicate noticeable solvent effects. This fact can be explained through different models, but independent of the chemistry of the system, to obtain the same copolymer composition in bulk or in solution, the monomer feed that should be used is determined by solvent polarity.  相似文献   

8.
The propagation rate coefficient kp for the free-radical polymerization of 1,3-butadiene was determined as a function of temperature in the range of 30–60°C in bulk and as a function of the concentration of 1,3-butadiene in chlorobenzene at 50°C using a combination of a UV flash-lamp induced polymerization and a measurement of the molecular weight distribution of the produced polymer by size exclusion chromamgraphy. Differences of up to around 45% resulted even for the same concentrations and temperatures when these measurements were compared with recent determinations of kp of 1,3-butadiene in chlorobenzene in which just such an experimental setup was used. This finding was essentially attributed to large differences in the radical termination rates which prevailed in the reaction volumes under the current conditions of the pulse-induced polymerization experiments. Moreover, in the vicinity of bulk concentrations of 1,3-butadiene in chlorobenzene, kp apparently decreases with concentration.  相似文献   

9.
The rate of polymerization (Rp) of 2-hydroxyethyl methacrylate, at low conversion, using aliphatic azo compounds as photoinitiators is considerably faster in water than in acetonitrile as a solvent. Active radical production in different solvents was analysed in detail. The quantum yield of active radicals showed only a minor dependence on the medium properties. These differences are due to changes in primary radical recombination reactions and they are unrelated to changes of Rp observed in different media. By employing intermittent illumination techniques, it is shown that the increased Rp in aqueous solution is due to a reduced rate of termination. Similar propagation rate constants were obtained in acetonitrile and water, whereas the termination rate constant increases by one order of magnitude when acetonitrile is employed instead of water. These results are discussed in terms of the effect of the cosolvent on the medium viscosity and the macroradical conformation.  相似文献   

10.
Pulsed-laser polymerization (PLP) in conjunction with molar mass distribution (MMD) measurement is the method of choice for determining the propagation rate coefficient kp in free-radical polymerizations. The authors, members of the IUPAC Working Party on Modeling of kinetics and processes of polymerization, collate results from using PLP-MMD to determine kp as a function of temperature T for bulk free-radical polymerization of methyl methacrylate at low conversions and ambient pressure. Despite coming from several different laboratories, the values of kp are in excellent agreement and obey consistency checks. These values are therefore recommended as constituting a benchmark data set, one that is best fitted by The 95% joint confidence interval for these Arrhenius parameters is also given. In so doing, we describe the most appropriate statistical methods for fitting kp(T) data and then obtaining a joint confidence interval for the fitted Arrhenius parameters. As well, we outline factors which impose slight limitations on the accuracy of the PLP-MMD technique for determining kp, factors which may apply even when this technique is functioning well. At the same time we discuss how such systematic errors in kp can be minimized.  相似文献   

11.
A methacrylamido-terminated saccharide, 6-deoxy-6-methacryloylamido-D -glucopyranose (MAG), was synthesized using a new and easy pathway. Kinetics of the radically initiated solution homopolymerization of this monomer was investigated at 50°C using 4,4′-azobis(4-cyanopentanoic acid); the obtained kp/kt1/2 value around 1.0 mol−1/2 ċ L1/2 ċ s−1/2 is indicative of the strong ability of MAG to polymerize (kp and kt being the rate constants of propagation and termination, respectively). The chemical structure of the homopolymer, as analyzed by 13C NMR spectroscopy shows a tacticity which is in favor of a syndiotactic configuration resulting from the bulkiness of the side group. The dimensional characteristics of the polymers were studied by light scattering as an on-line detection in size exclusion chromatography (SEC). As expected from the kp/kt1/2 value, high weight-average molecular weights were obtained. The small average radii of gyration were found to reflect a relatively compact structure in water due to inter- and intramolecular interactions; viscosity studies in various solvents as well as values of the second virial coefficient corroborate such a behavior which reflects that water is a poor solvent for poly(MAG).  相似文献   

12.
The rate of thermal decomposition of 1.4-dimethoxycarbonyl-1.4-diphenyl-2-tetrazene (DDT) has been measured in various solvents by UV spectroscopy. The rate measurements were carried out over a temperature interval from 65 to 85°C. The activation parameters have been calculated from the dependence of the rate constants on the temperature. On the basis of the kinetic data, the influence of the solvent on the decomposition of DDT was discussed. In this connection the solvent effect on the rate of polymerization of styrene initiated by DDT was also examined.  相似文献   

13.
The propagation rate coefficient Kp of the free-radical bulk polymerization of styrene is determined between 30 and 90°C up to a maximum pressure of 2800 bar. The data from pulsedlaser polymerizations and product analyses by gel-permeation chromatography (PLP-GPC) are adequately represented by the expression: The conseaquences of deducing activation volumes and activation energies from kp/(L · mol?1 · s?1) or fromk*p/(kg · mol-1 ·. S?1) are outlined.  相似文献   

14.
When a polymerizable system is subjected to periodic light flashes, which induce the formation of primary radicals, a pseudostationary state is established which is characterized by a periodic profile of the (polymer) radical concentration. Within such a period of length t0 the radical concentration will decay according to a second-order rate law. At the end of this period the radicals, which have escaped termination up to this moment, have propagated up to a chain length L0 = t0·kp·cM, kp representing the propagation rate constant and cM the monomer amount concentration. When the next flash arrives these radicals are opposed to a strongly increased overall concentration of radicals which leads to an enhanced probability for their termination. As a consequence the formation of dead polymer molecules with a chain length close to L0 is favoured. The chain-length distribution of polystyrene prepared under such pseudostationary conditions, which was evaluated by gel permeation chromatography, in fact exhibits such a peak. The analysis of the theoretical distribution curves, derived in this communication, reveals that it is easily possible to correlate this peak to L0, independently of the mode of termination (disproportionation or combination). Thus, a method of evaluating kp is derived without any reference to the termination rate constant kt and largely independent of all features which usually cause problems in the evaluation of kp and kt (such as primary radical termination etc.). The experimental results agree fairly well with the data reported in literature, especially with those obtained from the number of particles and the rate of polymerization in emulsion systems.  相似文献   

15.
Rate coefficients of termination and transfer in the free-radical polymerization of 1,3-butadiene in chlorobenzene were determined in the temperature range 318 K < T < 333 K. On the basis of an earlier published temperature dependence of the rate coefficient of propagation, for the termination reaction the Arrhenius equation Kt = 1,13 · 1010 · exp(? 711 K/T) L · mol?1 · s?1 was obtained. For the transfer to monomer the experiments yielded the Arrhenius equation Ktr,M = 4,22 · 106 · exp(? 5140 K/T) L · mol?1 · S?1 and for the transfer to the solvent Ktr,S = 2,25 · 108 · exp(? 7050 K/T) L · mol?1 · S?1.  相似文献   

16.
The condensation reaction of thioalanine S-dodecyl ester ( 1c ) was carried out in several organic solvents, or in their mixtures with water. In acetone, benzene, chloroform, or hexane, no cyclic dimer of alanine, 3,6-dimethyl-2,5-piperazinedione ( 2 ) and in the water/organic solvent mixtures only low yields of 2 were obtained, contrary to the good yields which resulted in water alone. In the condensation reaction of thioalanine S-octyl ester ( 1b ) obvious salt effects were observed, especially with FeCl3.  相似文献   

17.
Polymerization of bis(2-ethylhexyl) itaconate ( 1 ) with dimethyl azobis(isobutyrate) ( 2 ) was carried out at 50°C in various solvents. Polar solvents caused a significant decrease in the polymerization rate (Rp) and the molecular weight of resulting poly( 1 ). The propagating poly( 1 ) radical could be observed as a five-line ESR spectrum in the actual polymerization systems used. The stationary concentration of poly( 1 ) radical was determined by ESR to be 4,2–6,4 · 10?6 mol · L?1 at 50°C when the concentrations of 1 and 2 were 1,03 and 3,00 · 10?2 mol · L?1. Using Rp the monomer concentration and the polymer radical concentration, the propagation rate constant (Kp) was estimated to be 1,4–6,8 L · mol?1 · s?1, depending on the solvents used. The kp value was smaller in more polar solvents. The solvent effect is explained in terms of the solvent affinity for the propagating polymer chain.  相似文献   

18.
Vinyl trifluoroacetate was polymerized by free radicals in various kinds of solvents at temperatures between +60 and -40°C. The syndiotactic and isotactic diad contents of the polymer were determined via the triad content, which was obtained by their NMR-spectra. The differences of the activation enthalpies and those of the activation entropies between isotactic and syndiotactic arrangements were calculated from the temperature dependence of the ratio of isotactic to syndiotactic diads formed in the different solvents. A compensation effect exists between the two differences. The syndiotacticities of the polymers synthesized in alkanes in which the polymers were insoluble were higher than those in esters and ketones in which the polymers were soluble. The syndiotactic diad contents of the polymer synthesized in alkanes and ketones at ?40°C amount to 60~63% and 52~56% respectively.  相似文献   

19.
In order to investigate the initiation mechanism of the polymerization of MMA by N, N-dimethylaniline (DMA), p-substituted DMA's were prepared and used for a kinetic study. Electron donating groups increase and electron withdrawing substituents decrease the over-all rate of polymerization. Moreover, different kinds of amines were examined as initiators. The results were related with the Hammett equation and the ionization potential of the amines. The solvent effect and the composition of the initiator complex were investigated. From these results, it is concluded that the initiating species is a 1:1 electron transfer complex of DMA with the monomer methyl methacrylate.  相似文献   

20.
The living stereospecific polymerization of methylthiirane initiated by a chiral cadmium compound was run in different solvents. The general features of the mechanism of stereoregulation are the same as in bulk. In heptane and in heptane/propylene carbonate mixtures, the stereospecificity of the reaction does not depend on monomer concentration and on the dielectric constant ε of the medium. This independence of stereospecificity is in agreement with the absence of a variation of the complexation constant between a model cadmium derivative and the monomer with the variation of ε in a large range. The enantioasymmetric process depends on the molecular weight of the chain bearing the active sites. A change from diastereoisomeric to enantiomeric forms of active sites occurs for molecular weights around 50000. The enantioasymmetric constant depends on the nature of the medium and we observed a Born law for the dependence of the enantioasymmetric constant versus reciprocal dielectric constant: log kR ∝ 1/ε. This effect can be explained by a variation of the reactivities of the two types of active sites and occurs without a change in their relative amounts.  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