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.
A method is presented which allows the determination of kp/kt-values in free radical polymerization. It is based on measurements of the (average) rate of polymerization under pseudostationary conditions, the polymerization being initiated by laser flashes of short duration. For ρkt t0 ? 1 (ρ being the additional polymer radical concentration produced by each laser flash, kt the bimolecular termination constant between polymer radicals, kp the rate constant of chain propagation, t0 the time separating two successive laser flashes) kp/kt may be obtained as the slope of a linear plot of the fractional conversion per flash vs. ln t0. Dividing the intercept by the slope yields ln (pkt). Thus, if p is accessible, separation of kp/kt-data into its individual constituents may be accomplished without making any use of stationary polymerization data. Application of this method to the polymerization of styrene sensitized by benzoin or AIBN at 25°C gives kp/kt-values of 1,0 · 10?6 which are in fair agreement with those obtained by other methods.  相似文献   

2.
The free radical polymerization of butyl acrylate has been studied in benzene solutions ranging from 1 to 5 mol·L–1 at 50°C using 2,2′‐azobisisobutyronitrile as initiator. Under the conditions of our experiments, both the effective rate coefficient for initiation, 2 f kd , and the coupled parameter, kp/kt1/2, (where kp and kt are the constants for propagation and termination reactions, respectively) are dependent on the monomer concentration. The 2 f kd value shows little increase with monomer concentration. The variation of the kp/kt1/2 parameter has been correlated with the chain length dependence of the termination rate coefficient. This effect is also responsible for the high dependence of the overall polymerization rate, Rp, onthe monomer concentration (1.49).  相似文献   

3.
The polymerization of N-octadecylmaleimide ( 1 ) initiated with azodiisobutyronitrile ( 2 ) was investigated kinetically in benzene. The overall activation energy of the polymerization was calculated to be 94,2 kJ·mol?1. The polymerization rate (Rp) at 50°C is expressed by the equation, Rp = k[ 2 ]0,6[ 1 ]1,7. The homogeneous polymerization system involves ESR-detectable propagating polymer radicals. Using Rp and the polymer radical concentration determined by ESR, the rate constants of propagation (kp) and termination (kt) were evaluated at 50°C. kp (33 L · mol?1 · s?1 on the average) is substantially independent of the monomer concentration. On the other hand, kt (0,3 · 104 – 1,0 · 104 L · mol?1 · s?1) is fairly dependent on the monomer concentration, which is ascribable to a high dependence of kt on the chain length of rigid poly( 1 ). This is the predominant factor for the high order with respect to the monomer concentration in the rate equation. In the copolymerization of 1 (M1) and St (M2) with 2 in benzene at 50°C, the following copolymerization parameters were obtained: r1 = 0,11, r2 = 0,09, Q1 = 2,1, and e1 = +1,4.  相似文献   

4.
The single pulse (SP)‐pulsed‐laser polymerization (PLP) technique has been applied to measure kt/kp, the ratio of termination to propagation rate coefficients, for the free‐radical bulk polymerization of styrene at temperatures from 60 to 100°C and pressures from 1800 to 2 650 bar. kt/kp is obtained by fitting monomer concentration vs. time traces that are determined via time‐resolved (μs) near infrared monitoring of monomer conversion induced by single excimer laser pulses of about 20 ns width. Styrene is a difficult candidate for this kind of measurements as conversion per pulse is small for this low kp and high kt monomer. Thus between 160 to 300 SP signals were co‐added to yield a concentration vs. time trace of sufficient quality for deducing kt/kp with an accuracy of better than ± 20 per cent. With kp being known from PLP–SEC experiments, chain‐length averaged kt values are immediately obtained from kt/kp. At given pressure and temperature, kt is independent of the degree of overall monomer conversion, which, within the present study, has been as high as 20%percnt;. The kt value, however, is found to slightly increase with the amount of free radicals produced by a single pulse in laser‐induced decomposition of the photoinitiator DMPA (2,2‐dimethoxy‐2‐phenyl acetophenone). This remarkable observation is explained by DMPA decomposition resulting in the formation of two free radicals which significantly differ in reactivity. Extrapolation of SP–PLP kt data from experiments at rather different DMPA levels and laser pulse energies toward low primary free‐radical concentration, yields very satisfactory agreement of the extrapolated kt values with recent literature data from chemically and photochemically induced styrene polymerizations.  相似文献   

5.
The effect of SnCl4 on the polymerization of diethyl itaconate ( 1 ) with dimethyl 2,2′-azoisobutyrate ( 2 ) in benzene was investigated kinetically and ESR spectroscopically. The polymerization rate (Rp) at 50°C shows a flat maximum on varying the SnCl4 concentration. The molecular weight of the resulting polymer decreases with increasing SnCl4 concentration. The overall activation energy of the polymerization is lowered from 52 to 33 kJ · mol?1 by the presence of SnCl4, (0,342 mol · L?1). An NMR study revealed that 1 and SnCl4 form 1:1 and 2:1 complexes with a large stability constant in benzene. The propagating polymer radicals in the absence and presence of SnCl4 are ESR-observable as a five-line spectrum under the actual polymerization conditions. The complexed polymer radicals show further three-line splitting due to two methylene hydrogens of the ethyl ester group. The polymer radical concentration increases with the SnCl4 concentration. The rate constant (kp) of propagation was determined using Rp and the polymer radical concentration. kp (6,3–2,9 L · mol?1 · s?1 at 50°C) decreases with increasing SnCl4 concentration. The presence of SnCl4 (0,342 mol · L?1) reduces the activation energy of propagation from 29 to 21 kJ · mol?1. The rate constant (kt) of termination was estimated from the decay curve of the polymer radicals, kt (3,1–1,1 · 105 L · mol?1 s?1) also decreases with the SnCl4 concentration. The activation energies of termination in the absence and presence of SnCl4 (0,342 mol · L?1) are 30 and 24 kJ · mol?1, respectively. Suppression of propagation and termination by SnCl4 seems to be explicable in terms of an entropy factor.  相似文献   

6.
The low and intermediate conversion solution polymerization kinetics of methyl methacrylate in toluene and in 2-butanone are investigated. For this purpose, repetitively applied laser pulses are employed to stimulate initiation, and quantitative near infrared spectroscopy is used to determine methyl methacrylate concentrations. The monomer concentration change per pulse is measured; it is shown how this information may be converted into coupled values of the rate coefficient for propagation kp, the rate coefficient for termination kt, and the total efficiency of (photo)initiation. In the present case kp is already known from independent investigations, which enables our pseudostationary state methods to yield kt directly. The major thrust of this paper, therefore, concerns the effect of solvent on kt: it is found that at both low and intermediate conversions, kt increases as the solvent concentration increases; possible causes of this are discussed. The effect of initial solvent fraction on the variation of kt with conversion is also reported. Although all our experimental measurements are carried out at an elevated pressure (1000 bar) and a single temperature (30°C), there is no reason for suspecting that our findings regarding solution polymerization termination should not be representative for polymerizations of methyl methacrylate (and other methacrylate monomers) in general.  相似文献   

7.
The polymerization of methyl N-phenylitaconamate(methyl 2-methylenesuccinanilate ( 1 )) with dimethyl 2,2′-azodiisobutyrate ( 2 ) was studied in N,N-dimethylformamide (DMF) kinetically and by means of electron paramagnetic resonance (EPR) spectroscopy. The polymerization rate (Rp) at 55°C is given by the equation: Rp = k[ 2 ]0,58 · [ 1 ]1,6. The overall activation energy of the polymerization was calculated to be 54,2 kJ/mol. The number-average molecular weight of poly( 1 ) was in the range between 5000 and 17000. From an EPR study, the polymerization system was found to involve the EPR-detectable propagating polymer radical of 1 at practical polymerization conditions. Using the concentration of polymer radicals, the rate constants of propagation (kp) and termination (kt) were determined for 55°C. The rate constant of propagation kp (between 8,4 and 12 L · mol?1 · s?1) tends to somehow increase with increasing monomer concentration. On the other hand, kt (between 1,9. 10?5 L · mol?1 · s?1) increases with decreasing monomer concentration, which results from a considerable dependence of kt on the polymer-chain length. Such monomer-concentration-dependent kp and kt values are responsible for the high dependence of Rp on the monomer concentration. Thermogravimetric results showed that thermal degradation of poly( 1 ) occurs rapidly at temperatures higher than 200°C and the residue at 500°C amounts to 26% of the initial polymer. For the copolymerization of 1 (M1) with styrene (M2) at 55°C in DMF the following copolymerization parameters were found: r1 = 0,52, r2 = 0,31, and Q, e values Q1 = 1,09 and e1 = +0,55.  相似文献   

8.
The copolymerization of N-cyclohexylmaleimide ( 1 ) (M1) and bis(2-ethylhexyl) itaconate ( 2 ) (M2) with dimethyl 2,2′-azoisobutyrate ( 3 ) as an initiator was carried out at 50°C in benzene. Monomer reactivity ratios were estimated as r1 = 0,34 and r2 = 0,38. The copolymerization rate (Rp) and the molecular weight of the resulting copolymer increased with increasing concentration of 1 when the total concentration of comonomers was fixed at 1,00 mol. L?1. Rp was proportional to [ 3 ]0,5, indicating a usual bimolecular termination in the copolymerization. An electron spin resonance (ESR) spectrum of the propagating polymer radicals was observable in the actual copolymerization system at 50°C. The spectrum of the copolymerization system is inexplicable in terms of any superposition of spectra observed in the corresponding homopolymerization systems, revealing that some penultimate monomeric unit causes a change in the ESR spectrum, that is, the structure of propagating polymer radical. The apparent rate constant of propagation (kp) and termination (kt) were estimated by ESR. The kp values (1,5–50 L · mol?1 · s?1) are fairly higher than those estimated on the basis of the terminal model, affording another piece of evidence for the penultimate effect. The kt value (1,8–5,4·103 L · mol?1 · s?1) shows a behaviour similar to that of the intrinsic viscosity of the resulting copolymer on varying the monomer feed composition, which seems to reflect diffusion-control of termination reactions.  相似文献   

9.
In this work the exact data of the rate coefficients of propagation kp and termination kt, and initiator efficiency f as well as their variations with conversion during the whole processes of the polymerization of methyl methacrylate in bulk have been measured using electron paramagnetic resonance (ESR) without any assumptions and approximations. kp was calculated directly from the synchronously measured data of the concentration of radicals and polymerization rate. kt was determined under nonsteady-state conditions by using the after-effect technique in the ESR measurement. f for the initiation with dimethyl 2,2′-azoisobutyrate was evaluated from the initiation and termination rates. The ESR spectra show that the polymerization system is in a micro-heterogeneous state and there exist inactive radicals. A model of a diffusion-controlled reaction is proposed.  相似文献   

10.
We describe our method based on pulse radiolysis with optical detection developed for the examination of the kinetics and mechanism of the first steps of high‐energy radiation initiated polymerization. The absorption spectra of the intermediates were obtained in cyclohexane solution of hexanediol diacrylate (HDDA) of different concentrations. In dilute solution (10 mmol·dm–3) and short time (10 μs) after the pulse, the spectrum of the monomer radicals was observed. On increasing the monomer concentration, the maximum of the spectrum was shifted to longer wavelength indicating the start of the oligomerization reaction. The increase in the time of observation resulted in a similar shift in dilute solution. From the kinetic curves the rate coefficients of termination for the monomer radicals (2·kt,m) and average rate coefficients of termination for the oligomer radicals of different chain length (2·kt) were determined. The average rate coefficient of termination was found to decrease in time (that is with increasing chain‐length).  相似文献   

11.
The absolute rate constants of propagation kp and of termination kt of ethyl α-cyanoacrylate (ECNA) were determined in bulk at 30°C by means of the rotating sector method under conditions to suppress anionic polymerization; kp = 1 622 1 · mol?1 · s?1 and kt = 4,11 · 108 1 · mol?1 · s?1 for the polymerization in the presence of acetic acid, and kp = 1610 1 · mol?1 · s?1 and kt = 4,04 · 108 l · mol?1 · s?1 for the polymerization in the presence of 1,3-propanesultone. The magnitude of k/kt determined was 6,39 · 10?3 l · mol?1 · s?1. The absolute rate constants for cross-propagation in ECNA copolymerizations were also evaluated. Quantitative comparison of the rate constants with those of common monomers and polymer radicals shows that the strong electron-withdrawing power of the ethoxycarbonyl and cyano groups enable the poly(ECNA) radical to add to monomers as fast as the other polymer radicals. The relatively high reactivity of ECNA, regardless of the type of attacking polymer radical, is interpreted by a transition state greatly stabilized by both the ethoxycarbonyl and the cyano groups.  相似文献   

12.
The polymerization of N-(2,6-dimethylphenyl)itaconimide (1) with azoisobutyronitrile (2) was studied in tetrahydrofuran (THF) kinetically and spectroscopically with the electron spin resonance (ESR) method. The polymerization rate (Rp) at 50°C is given by the equation: Rp = K [2] 0,5 · [1] 2,1. The overall activation energy of the polymerization was calculated to be 91 kJ/mol. The number-average molecular weight of poly (1) was in the range of 3500–6500. From an ESR study, the polymerization system was found to involve ESR-observable propagating polymer radicals of 1 under the actual polymerization conditions. Using the polymer radical concentration, the rate constants of propagation (kp) and termination (kt) were determined at 50°C. kp (24–27 L · mol?1 · s?1) is almost independent of monomer concentration. On the other hand, kt (3,8 · 104–2,0 · 105 L · mol?1 · s?1) increases with decreasing monomer concentration, which seems mainly responsible for the high dependence of Rp on monomer concentration. Thermogravimetric results showed that thermal degradation of poly (1) occurs rapidly at temperatures higher than 360°C and the residue at 500°C was 12% of the initial polymer. For the copolymerization of 1 (M1) with styrene (M2) at 50°C in THF the following copolymerization parameters were found; r1 = 0,29, r2 = 0,08, Q1 = 2,6, and e1 = +1,1.  相似文献   

13.
The propagation rate coefficient, kp, of poly(ethylene glycol) methyl ether methacrylate (Mn ≈500 g mol?1) has been measured via pulsed‐laser polymerization (PLP)–size‐exclusion‐chromatography in aqueous solution between 5 wt% monomer and bulk polymerization at temperatures from 22 to 77 °C. kp increases significantly toward higher water content, as is observed for other water‐soluble monomers. This entropy‐motivated effect enhances the pre‐exponential. The activation energy, EA(kp), is more or less identical to the characteristic value of methacrylates. The chain‐length‐dependent rate coefficient, kti,i, for termination of two radicals of chain length i has been investigated at low degrees of monomer conversion via the single‐pulse–PLP–electron paramagnetic resonance technique. kti,i turned out to be adequately represented by the composite model designed by the Russell group. The power‐law exponents for the chain‐length dependence of small and long radicals are close to the numbers reported for other monomers. The rate coefficient for termination of two radicals of chain length unity scales with the fluidity of the reaction mixture. Viscosity measurements prior to polymerization thus enable estimates of termination rate.

  相似文献   


14.
Polymerization of dialkyl itaconates with dimethyl azoisobutyrate ( 5 ) was studied in benzene at 50°C by means of electron spin resonance (ESR). The monomers used are dimethyl ( 1 ), diethyl ( 2 ), dibutyl ( 3 ) and di-2-ethylhexyl ( 4 ) itaconates. All the polymerization systems involve ESR-observable propagating polymer radicals under the actual polymerization conditions. The polymerization rate (Rp) and degree of polymerization of the resulting polymer increase in going from shorter to longer alkyl groups. The ESR-determined rate constants of propagation (kp) and termination (kt) decrease as the alkyl chain becomes longer. kp of 1 is 3,3 times higher than that of 4 , while kt of 1 is 590 times higher than that of 4 . Thus, the steric effect due to the alkyl group suppresses much more termination than propagation, leading to the fact that Rp increases as the alkyl group becomes larger.  相似文献   

15.
The termination kinetics of the free‐radical bulk copolymerization of dodecyl acrylate (DA) and methyl acrylate (MA) has been investigated at various monomer mole fractions between 15 and 50 °C and up to 2000 bar. The ratio of termination to propagation rate coefficients, (kt/kp)copo, is measured via the single pulse‐pulsed laser polymerization (SP‐PLP) technique. Chain‐length averaged kt,copo are deduced from (kt/kp)copo in conjunction with kp,copo data that are estimated via a simplified version of the implicit penultimate unit effect (IPUE) model. At low and moderate degrees of monomer conversion extended ranges of almost constant kt are observed where termination is controlled by segmental diffusion with important contributions of steric effects. These plateau kt values are significantly – by almost two orders of magnitude – different for MA and DA. The increase with MA content of kt,copo is adequately described by a penultimate unit model which uses the geometric mean approximation for estimating rate coefficients of cross‐termination between radicals of different free‐radical terminus. The model applies within the entire pressure and temperature range of the present study. At high degrees of monomer conversion, at and above 50%, homo‐kt and kt,copo are almost insensitive toward the composition of the monomer mixture. The termination rate under these conditions is essentially controlled by reaction diffusion.

Relative monomer conversion, cM(t)/cM,0, vs. time, t, plot of a methyl acrylate (MA) – dodecyl acrylate (DA) copolymerization (fMA,0 = 0.5) induced by a single laser pulse at 30 °C and 10 bar. The difference between measured and fitted (to Equation (4)) data is illustrated by the plot of residuals (res) in the lower part of the figure.  相似文献   


16.
Pulsed laser photolysis (PLP) has been employed to determine propagation rate constants kp for styrene polymerization in benzene over a wider temperature range (20?80°C) than previously converd. It is proposed that a small chain length dependence of kp (overall) may, in part, be a consequence of a marked chain length dependence of kp for the first few propagation steps [i.e. kp(1) > kp(2) < kp(3) ≥ kp(≥4)]. The propagation rate constant for styrene polymerization is given by the expression: In kp = 16,09 ? 28950/(RT) (overall) or In kp = 16,47 ? 30084/(RT) (chain length ≥ 4). Kinetic simulation has been applied both as an aid in data analysis and to demonstrate the reliability of the PLP technique for evaluation of propagation rate constants (kp) in radical polymerization. This has been achieved by examining the sensitivity of the molecular weight distribution of polymers formed in PLP experiments to the values of the kinetic parameters associated with polymerization and their chain length dependence. The termination rate constants (kt = kc + kd) and the ratio of combination to disproportionation (kc: kd) markedly affect the molecular weight distribution of polymer formed in PLP experiments. The prospects for evaluating the values of kt, its chain length dependence and kc : kd by direct analysis of the molecular weight distribution are discussed in the light of these results.  相似文献   

17.
In order to assess the general consequences of an eventual chain length dependence of the bimolecular termination rate constant kt in free radical polymerization, an iterative procedure for solving the kinetic scheme was developed for this case. The iteration process is based on the equation for the propagation probability of a radical chain of length x, αx, which constitutes a cyclic definition of the α-values, with F(x,y) giving the functional dependence of kt = k F(x, y) on the lengths x and y of the two radicals involved in the termination process. As, in addition, αx links the stationary concentrations of radicals of adjacent length, [Rx?1] and [Rx], solving for the set of αx allows to build up the complete chain length distribution of polymer radicals from which all relevant kinetic quantities may be calculated. The following types of functional dependence have been used in practice The case of F2, which is the slightly less realistic one, allows to separate the variables x and y in the equation for αx, thus leading to an immediate comfortable solution. Irrespective of the type of F(x, y) chosen (the differences being quantitative in nature rather than qualitative) it can be shown that nearly all the familiar relationships known from “classical” polymerization kinetics with chain length independent termination (b = 0) break down. The new relationships, however, can be cast into comparatively simple power laws. The general consequences of the concept of chain length dependent termination are discussed in some detail.  相似文献   

18.
The free radical polymerisation kinetics of diethyl- (DEI), dipropyl- (DnPI), dibutyl- (DnBI), and dihexyl itaconate (DHI) in the bulk were studied in the temperature range from 50 to 70°C. The concentration of the initiator, 2,2′-azoisobutyronitrile (AIBN), was varied between 0.02 and 0.085 mol/dm3. The rate of polymerisation (Rp), degree of polymerisation (DP), overall polymerization rate constant (K), the ratio of the propagation and termination rate constants (kp/kt1/2), as well as the chain transfer constant to monomer (CM) were determined. The values of Rp, K, and kp/kt1/2 of the investigated monomers all increase with increasing size of the alkyl group in the ester substituent, whereas CM decreases when going from the dimethyl to the dihexyl ester. The values of CM are larger than the corresponding values for the alkyl esters of methacrylic acid.  相似文献   

19.
The homogeneous and the dispersion polymerzation and copolymerization of methacryloyl-terminated poly(oxyethylene) (MMA-PEG) and of p-vinylbenzyl-terminated poly(oxyethylene) (St-PEG) macromonomers and styrene, initiated by a radical initiator, was investigated using conventional gravimetric and NMR methods at 60°C. The batch polymerizations in N,N-di-methylformamide and in ethanol/water were conducted to either low or high conversion. The fractional conversion rates of the solution polymerization and copolymerization indicate that the homopolymerizations of macromonomers involve steady-state conditions, whereas copolymerizations proceed under non-stationary conditions. The ratios of the rate constants for propagation and termination (kp/k) for polymerization and copolymerization of MMA-PEG and St-PEG are by one order of magnitude higher than that for styrene. The increase in kp/k is more pronounced in dispersion polymerization, which is ascribed to the decrease of both kp and kt. The rates of dispersion polymerization are proportional to the particle concentration. The number of particles increses up to 50% conversion. The particle growth is suggested to proceed via association of partcles and by propagation within polymer particles. The decrese of the number of radicals per particles as conversion proceeds is ascribed to the decrese of the growing radical activity and to the transfer of monomeric radicals to the continuous phase. The molecular weights correlate inversely with the particle size.  相似文献   

20.
Phenethyl 2-(methoxycarbonylmethyl)acrylate (methyl penethyl itaconate) ( 1 ) was prepared and its polymerization with dimethyl 2,2′-azoisobutyrate ( 2 ) was investigated kinetically in benzene. The polymerization rate (Rp) was found to be expressed by Rp = k·[ 2 ]0,5·[ 1 ]1,8 (at 50°C). Further, a higher dependence of Rp (2nd order) on the concentration of 1 was observed at 61°C. The overall activation energy of the polymerization was calculated to be a low value of 50,3 kJ/mol. The initiator efficiency (f) of 2 was determined to be 0,48 to 0,22 at 50°Cand 0,50 to 0,28 at 61°C. f decreases with increasing monomer concentration due to the high viscosity of 1 . The poly( 1 ) radical is stable enough to be observable by ESR at high temperatures (50–60°C). Rate constants of propagation (kp) and termination (kt) were estimated using the poly( 1 ) radical concentration determined by ESR. kp [6,0 to 121/(mol·s) at 50°Cand 7,1 to 15 1/(mol·s) at 61°C] shows a trend to increase with the concentration of 1 . On the other hand, kt [2,9·104 to 17·1041/(mol·s) at 50°Cand 6,9·104 to 45·1041/(mol. s) at 61°C] decreases with increasing MPI concentration. This behavior is responsible for the high order with respect to monomer concentration. Copolymerization of 1 (M1) with styrene (M2) at 50°Cgave the following results: r1 = 0,36, r2 = 0,40, Q1 = 0,82 and e1 = + 0,59. Using the above results, the rate constants of the cross-propagations were estimated to be k12 = 22 1/(mol·s) and k21 = 308 1/(mol·s) at 50°C.  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