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.
The bulk polymerization of 2‐ethylhexyl acrylate (2‐EHA), induced by a pulsed electron beam, was investigated with pulse radiolysis, gravimetry, and Fourier transform infrared spectroscopy. The roles of the dose rate, pulse frequency, and added acrylic acid (AA) in the polymerization of 2‐EHA were examined at ambient temperature. In the range of 12.6–71.2 Gy/pulse, the polymerization of 2‐EHA was dose‐rate‐dependent: at the same total dose, a lower dose rate yielded a higher conversion. Also, a lower pulse rate gave a higher conversion at the same total dose. The addition of up to 10 wt % AA showed no increase in the conversion of 2‐EHA at a low conversion (8 kGy), but at a higher conversion (16 kGy), a 20 wt % increase in the conversion of 2‐EHA was observed. The estimated values (1.6 ± 0.3) × 10?3 (dm3 s)3/2 mol?1 s?1/2 for kp(G/2kt)1/2 and 2.6 ± 0.8 dm3 s J?1 for 2ktG (where kp is the rate constant of propagation, kt is the rate constant of bimolecular termination, and G is the yield of free radicals) were obtained at relatively low conversions. The reaction rate constant of the addition of 2‐EHA· free radicals to the monomer was measured by pulse radiolysis and found to be 2.8 × 102 mol?1 dm3 s?1. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 196–203, 2003  相似文献   

2.
The effect of a range of 10 organic nitriles on the radiation-induced polymerization of styrene was studied. A dose rate of 4.4 rad s?1 was used. A rate of polymerization of styrene (1.744 mol L?1 of toluene solution) of 5.0 × 10?7 mol L?1 s?1 was found. With organic nitriles present (styrene:nitrile ratio of 1:0.28) the rate of polymerization increased. Rates in the range of 5.5 × 10?7 ?5.2 × 10?6 mol L?1 s?1, depending on the nitrile present, were obtained. The polymers were partially characterized and evidence of involvement of each of the nitriles in the polymer chains was revealed. The increase in rate of polymerization has been attributed to the part played by nitrile radicals in the initiation of styrene polymerization. Radical yield values [as G(nitrile radical)] were derived from the relevant rate expressions. Values ranged from 2.7 to 49.5, depending on the particular nitrile. Corresponding values of G(nitrile radical) in the range of 5.1–129.4 were obtained by the manipulation of number-average molar mass data. Values of kpkt of approximately 2 × 10?5 L mol?1 s?1 were found. Trommsdorff types of effect are absent from these systems.  相似文献   

3.

Electrochemical polymerization of azure B from sulfuric acid solution was carried out by using cyclic voltammetry. The electrolytic solution consisted of 5.0 mmol · dm?3 azure B and 0.3 mol · dm?3 H2SO4. The temperature for polymerization was controlled at 20°C. A blue film, i.e., poly(azure B) was formed on a platinum foil and had a electrochemical reversibility, stability and a fast charge transfer ability in the 0.5 mol · dm?3 Na2SO4 with pH ≤4.0 solution. The currents of both anodic and cathodic peaks are proportional to υ1/2 at the scan rate (υ) region of 25 and 600 mV · s?1 on the cyclic voltammograms. The conductivity of poly(azure B) is 2.8×10?6 S · cm?1 at 20°C. The UV‐visible spectrum and Raman spectrum of the polymer are different from those of the monomer. A possible polymerization mechanism of azure B was also proposed.  相似文献   

4.
The spectroscopic and kinetic data of the short lived intermediates obtained by the attack of H-radicals on fluoro-, chloro-, bromobenzene, benzylchloride and phenethylchloride in aqueous solutions were studied by pulse radiolysis technique. The first three yield cyclohexadienylradicals (k=1–1.5×109 dm3 mol?1 s?1) with characteristic absorption maxima in the region 220–330 nm. In the case of benzylchloride a quantitative abstraction of chlorine by the H-atoms is observed (k=9.5×108 dm3 mol?1 s?1) leading to the formation of the benzylradical (λmax=257, 303, 317.5nm). The attack of H-atoms on phenethylchloride can occur on the aromatic ring forming also a cyclohexadienylradical (k=2.0×109 dm3 mol?1 s?1, λmax=317, 323nm) as well as on the side chain (k=1.5×108 dm3 mol?1 s?1) yielding H2. The intermediates decay according to a second order reaction withk=2 to 4.6×109 dm3 mol?1 s?1. To elucidate reaction mechanisms, steady state radiolysis experiments on the same systems were performed.  相似文献   

5.
A simplified kinetic model for RAFT microemulsion polymerization has been developed to facilitate the investigation of the effects of slow fragmentation of the intermediate macro‐RAFT radical, termination reactions, and diffusion rate of the chain transfer agent to the locus of polymerization on the control of the polymerization and the rate of monomer conversion. This simplified model captures the experimentally observed decrease in the rate of polymerization, and the shift of the rate maximum to conversions less than the 39% conversion predicted by the Morgan model for uncontrolled microemulsion polymerizations. The model shows that the short, but finite, lifetime of the intermediate macro‐RAFT radical (1.3 × 10?4–1.3 × 10?2 s) causes the observed rate retardation in RAFT microemulsion polymerizations of butyl acrylate with the chain transfer agent methyl‐2‐(O‐ethylxanthyl)propionate. The calculated magnitude of the fragmentation rate constant (kf = 4.0 × 101–4.0 × 103 s?1) is greater than the literature values for bulk RAFT polymerizations that only consider slow fragmentation of the macro‐RAFT radical and not termination (kf = 10?2 s?1). This is consistent with the finding that slow fragmentation promotes biradical termination in RAFT microemulsion polymerizations. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 604–613, 2010  相似文献   

6.
The free‐radical copolymerization of m‐isopropenyl‐α,α′‐dimethylbenzyl isocyanate (TMI) and styrene was studied with 1H NMR kinetic experiments at 70 °C. Monomer conversion vs time data were used to determine the ratio kp × kt?0.5 for various comonomer mixture compositions (where kp is the propagation rate coefficient and kt is the termination rate coefficient). The ratio kp × kt?0.5 varied from 25.9 × 10?3 L0.5 mol?0.5 s?0.5 for pure styrene to 2.03 × 10?3 L0.5 mol?0.5 s?0.5 for 73 mol % TMI, indicating a significant decrease in the rate of polymerization with increasing TMI content in the reaction mixture. Traces of the individual monomer conversion versus time were used to map out the comonomer mixture composition drift up to overall monomer conversions of 35%. Within this conversion range, a slight but significant depletion of styrene in the monomer feed was observed. This depletion became more pronounced at higher levels of TMI in the initial comonomer mixture. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 1064–1074, 2002  相似文献   

7.
Abstract

The balance between kinetics and thermodynamics is illustrated herein by the first direct polymerization of vinyl alcohol, the thermodynamically unstable tautomer of acetaldehyde, at a rate faster than it can tautomerize. Vinyl alcohol was formed through the acid catalyzed hydrolysis of ketene methyl vinyl acetal. With excess water present, the kinetics of tautomerization first order dependence upon vinyl alcohol (kobs = 2.73 × 10?4 s?1). Under water starved conditions, however, the kinetics now show a zero order dependence upon the concentration of vinyl alcohol (kobs = 3.5 × 10?6 M/s). Under these latter conditions, the half life of vinyl alcohol is nearly 24 hours at room temperature. Although cationic and homo free radical polymerization of vinyl alcohol failed, we found that this meta-stable species could be quantitatively polymerized in a copolymerization (AIBN, hυ, -10 to 25°C) with maleic anhydride. The kobs for copolymerization was found to be 4.41 × 10?4 sec?1 at ?10°C. Since the rate of polymerization is far greater than that of tautomerization under these conditions (ca. 30 times faster at ?10°C), there is no significant increase in acetaldehyde concentration during polymerization.  相似文献   

8.
The oxidation processes of the radiation-generated, three-electron-bonded intermediates AcMet2 [S??S]+ and AcMet [S??Br] were investigated by pulse radiolysis via their reactions with tryptophan (TrpH). These intermediates were derived from N-acetyl-methionine amide (N-AcMetNH2) and N-acetyl-methionine methyl ester (N-AcMetOMe). The bimolecular rate constant k of the reaction between each intermediate and l-tryptophan (TrpH) was measured. For N-AcMetNH2, k for the reaction of AcMet2 [S??S]+ with TrpH were 3.4?×?108 and 2.2?×?108?dm3?mol?1?s?1 at pH?=?1 and 4.5, respectively. For N-AcMetOMe, k for the reaction of AcMet2 [S??S]+ with TrpH were 4.0?×?108 and 2.8?×?108?dm3?mol?1?s?1 at pH 1 and 4.5, respectively. The rate constants for the intermolecular transformation of Met [S??Br] into TrpH+ or Trp were also estimated. For N-AcMetNH2, k for the reaction of AcMet2 [S??Br] with TrpH were 2.6?×?108 and 3.3?×?108?dm3?mol?1?s?1 at pH 1 and 4.5, respectively. Related mechanisms were discussed.  相似文献   

9.
The concentration of water in purified and BaO-dried α-methylstyrene was found to be 1.1 × 10?4M. The radiation-induced bulk polymerization of the α-methylstyrene thus prepared was studied in the temperature range of ?20°C to 35°C. The polymerization rate varied as the 0.55 power of the dose rate. The theoretical molecular weights and molecular weight distribution were calculated from a proposed kinetic scheme and these values were then compared with those found experimentally. The agreement between these two was reasonably close, and therefore it was concluded that, from the molecular weight distribution point of view, the proposed kinetic scheme for the cationic polymerization of α-methylstyrene is an acceptable one. The rate constant for chain transfer to monomer kf changed with temperature and was found to be responsible for the decrease in the molecular weight of the polymer with increase in temperature. kf and kp at 20°C were found to be 0.95 × 104 l./mole-sec and 0.99 × 106 l./mole-sec, respectively.  相似文献   

10.
The spectrocoulometric technique reported earlier is applied to verify the mechanism and to evaluate the contributions kBi of the individual bases to the total rate constant k of the hydrolysis of the tris (1,10-phenanthroline) iron(III) complex, Fe (phen)3+3. Both normal and “open-circuit” spectrocoulometric experiments are used. Partial rate constants for four bases in the acetate-buffered solutions are kH2O=(3.4±1.2) × 10?4s?1 (kH2O includes the H2O concentration), kOH=(1.20±0.06)×107 mol?1dm3s?1, kphen=(1.4±0.2) mol?1dm3s?1, kAc=(3.8±0.3)×10?2 mol?1dm3s?1, at 25°C and ionic strength 0.5 mol dm?3. The Fe(phen)3+3 hydrolysis, with (phen)2 (H2O) Fe-O-Fe (H2O) (phen)4+2 formation, is first order with respect to Fe (phen)3+3 and the bases present in the solution. The rate-determining step in the hydrolysis is the entry of a base to the coordinating sphere of the complex, as in the hydrolysis of the analogous 2,2′-bipyridyl complex.  相似文献   

11.
The kinetics of acrylamide polymerization has been investigated by employing cericammoniumnitrate-2-chloroethanol redox pair under nitrogen atmosphere at 30 ± 1°C. The rate of monomer disappearance is directly proportional to the concentration of 2-chloroethanol (1.0 × 10?2 ? 10.0 × 10?2 mol. dm?3) and is inversely proportional to the ceric ion concentration (2.5 × 10?3 ? 10.0 × 10?3 mol. dm?3) but shows square dependence to the concentration of monomer (5.0 × 10?2 ? 25.0 × 10?2 mol. dm?3). The rate of ceric ion disappearance is directly proportional to the initial concentration of ceric ion and 2-chloroethanol but independent of acrylamide concentration. The viscometric average molecular weight (M v) decreases on increasing the concentration of ceric ion and increases on increasing the concentrations of acrylamide and 2-chloroethanol. A tentative mechanism has been proposed.  相似文献   

12.
2-Vinyl anthraquinone has been polymerized, via radiation-induced initiation, in dimethyl-sulphoxide and in dichloromethane. Solvent to monomer ratios of 1 : 0.030 to 1 : 0.0167 mol have been examined for dose rates in the range 0.035 to 0.129 Gy s?1 and exposure times in the range 1.44 × 104 to 4.32 × 106 s. Rates of polymerization were found to lie in the range from 2.4 × 10?8 to 1.92 × 10?6 mol L?1 s?1. Values for the chain transfer constant to dimethyl-sulphoxide and to dichloromethane have been calculated. In addition, values of the kinetic ratio k/kt, for the polymerization of 2-vinyl anthraquinone have been estimated. The dependence of Rp on the monomer concentration and on the radiation intensity have been shown to be in broad agreement with a simplified steady-state kinetic scheme. A value of G(radical) for 2-vinyl anthraquinone was obtained via electron spin resonance studies, and shown to be 0.24. This G(radical) value and those obtained indirectly from kinetic data are discussed in relation to the molar mass values obtained from the poly(2-vinyl anthraquinone) products.  相似文献   

13.
The kinetics of the reaction of OH radicals with methyl, n-propyl, and n-butyl nitrite have been studied in a discharge flow system under pseudo first-order conditions. The OH radicals were generated by the reaction of H atoms with NO2 and the concentration of OH; monitored by resonance fluorescence, was followed as a function of time in an excess of each nitrite. Values of k(CH3ONO) = (0.6 ± 0.09) × 109 dm3 mol?1 s?1 k(n – C3H7ONO) = (1.39 ± 0.20) × 109 dm3 mol?1 s?1, and k(n – C4H9ONO) = (2.89 ± 0.43) × 109 dm3 mol?1 s?1 at 295 K were obtained. These results agree with previous relative rate measurements from this laboratory but the value for k (CH3ONO) is a factor of 7 greater than the value obtained by relative rate measurements elsewhere using a different OH source.  相似文献   

14.
Abstract

In this work was evaluated the activity of samarium acetate (III) (Sm(OAc)3) as a possible initiator in the polymerization by ring opening of trimethylene carbonate (TMC). All polymerizations were carried out under solvent-free melt conditions in ampoules-like flasks, equipped with a magnetic stirrer. The effects of different parameters of reaction, such as molar ratio monomer to initiator, temperature and reaction time, on typical variables of polymers, e.g., conversion of TMC to poly(trimethylene carbonate) (PTMC), dispersity and molar mass, were analyzed. The molar ratio of monomer to initiator was varied between 0 and 1000?mol/mol and the temperature among 70 and 150?°C. Nuclear Magnetic Resonance (1H-NMR and HMBC) and Size Exclusion Chromatography (SEC) were used to characterize the polymers. The results indicate that the Sm(OAc)3 induces the polymerization of TMC to high conversion with number-average molecular weights of 3.11?×?103 to 38.40?×?103?Da. Based on the 1H-NMR end-group analysis of low-molecular-weight PTMC, it was proposed a coordination–insertion mechanism for the polymerization, with a breakdown of the acyl-oxygen bond of the TMC. In according to the kinetic study carried out, the polymerization rate is first-order with respect to monomer concentration with apparent rate constants of kap?=?7.02?×?10?4?mol?×?L?1?×?h?1.  相似文献   

15.
The 1,1‐diphenylethene (DPE) controlled radical polymerization of methyl methacrylate was performed at 80 °C by using AIBN as an initiator and DPE as a control agent. It was found that the molecular weight of polymer remained constant with monomer conversion throughout the polymerization regardless of the amounts of DPE and initiator in formulation. To understand the result of constant molecular weight of living polymers in DPE controlled radical polymerization, a living kinetic model was established in this research to evaluate all the rate constants involved in the DPE mechanism. The rate constant k2, corresponding to the reactivation reaction of the DPE capped dormant chains, was found to be very small at 80 °C (1 × 10?5 s?1), that accounted for the result of constant molecular weight of polymers throughout the polymerization, analogous to a traditional free radical polymerization system that polymer chains were terminated by chain transfer. The polydispersity index (PDI) of living polymers was well controlled <1.5. The low PDI of obtained living polymers was due to the fact that the rate of growing chains capped by DPE was comparable with the rate of propagation. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2009  相似文献   

16.
Kinetics of the photoaquation of hexacyanoferrate(II) ion in aqueous solution were studied potentiometrically and spectrophotometrically. Supposing the simplest mechanism (see Fig. 3. in text), the photoaquation in alkaline medium can be well described. The value of the constants at pH = ll.0 are: ø = 0.8-1.0, k6 = (3.0 ± 0.5) × 10?8 s?1 and k?6 = 1.5 ± 0.2 mol?1 dm3 s?1. To describe the photoaquation in neutral medium t was extended (k′ = 3.33 x 102 mol?1 dm3s?1). The quantum yield in acidic medium can be calculated by combination of ø values of different protonated complexes. The reversibility of photoaquation in alkaline medium is also explained by the scheme.  相似文献   

17.
《Electroanalysis》2004,16(20):1690-1696
The electrode mechanism of Mo(VI) reduction was studied under catalytic adsorptive stripping mode by means of square‐wave voltammetry (SWV). Mo(VI) creates a stable surface active complex with mandelic acid. The electrode reaction of Mo(VI)‐mandelic acid system undergoes as one‐electron reduction, exhibiting properties of a surface electrode process. In the presence of chlorate, bromate, and hydrogen peroxide, the electrode reaction is transposed into a catalytic mechanism. The experimental results are compared with the recent theory for surface catalytic reaction, enabling qualitative characterization of the electrode mechanism in the presence of different catalytic agents. Utilizing both the method of “split SW peaks” and “quasireversible maximum” the standard redox rate constant of Mo(VI)‐mandelic acid system was estimates as ks=150±5 s?1. By fitting the experimental and theoretical results, the following catalytic rate constants have been estimated: (8.0±0.5)×104 mol?1 dm3 s?1, (1.0±0.1)×105 mol?1 dm3 s?1, and (3.2±0.1)×106 mol?1 dm3 s?1, for hydrogen peroxide, chlorate, and bromate, respectively.  相似文献   

18.
Decene-l was polymerized with the MgCl2/ethylebenzoate/p-cresol/AIEt3/TiCl4-AlEt3/methyl-p-toluate catalyst at 50° using an A/T ratio of 167 and a range of monomer concentration. The concentration of the two kinds of active sites are [Ti] = 12% and [Ti] = 4% of the total titanium. The rate constants of propagation are 24 M?1 s?1. Chain transfers to AIEt3, monomer, and by β-hydride elimination have rate constant values of 1.7 × 10?3 M?1 s?1, 1.34 × 10?2 M?1 s?1, and 1.7 × 10?2 s?1, respectively. Poly(decene-l) have relatively narrow MW which are unchanged during the course of a polymerization. Therefore, the active site concentrations in the CW catalyst for propylene and decene polymerization are identical and their rate constant values agree within a factor of 2. However, the rate of decene polymerization depends on fractional order of monomer concentration and decreases with the increase of activator concentration. Furthermore, the formation of metal polymer bonds has a rate independent of these concentrations. These kinetic behaviors are a manifestation of absorption processes of these species which are not seen in propylene polymerizations.  相似文献   

19.
The polymerization of diallyl phthalate has been studied in two solvents, benzene (GRadical = 0.7) and chloroform (GR = 11.2), γ-radiation being used to investigate the effect of the solvent on the rates of polymerization and also chain transfer to the solvent. Kinetic analysis shows that in benzene solution the initiating species come almost exclusively from the monomer, but in chloroform they arise only from the solvent. The latter was further confirmed from the chlorine analysis of the polymer wherein chloroform appears to have telomerized with diallyl phthalate. In neither of the solvents was high molecular weight polymer obtained. The kp/kt1/2 for the polymerization of DAP was found to be 3.3 × 10?4 and 1.17 × 10?3 in benzene and chloroform solutions, respectively. The chain-transfer constant CS was 11.25 × 10?3 and 9.75 × 10?3 for benzene and chloroform, respectively.  相似文献   

20.
The kinetics of ethylene/propylene copolymerization catalyzed by (ethylene bis (indeyl)-ZrCI2/methylaluminoxane) has been investigated. Radiolabeling found about 80% of the Zr to be catalytically active. The estimates for rate constants at 50°C are k11 = 1104 (Ms)?1, k12 = 430 (Ms)?1, k22 = 396 (Ms)?1,k21 = 1020 (Ms)?1, and kAtr,1 + kAtr.2 = 1.9 × 10?3 s?1. Substitution of trimethylaluminum for methylaluminoxane resulted in proportionate decrease in polymerization rate. The molecular weight of the copolymer is slightly increased by loweing the [Al]/[Zr] ratio, or addition of Lewis base modifier but at the expense of lowered catalytic activity and increase in ethylene content in the copolymer. Lowering of the polymerization temperature to 0°C resulted in a doubling of molecular weight but suffered 10-fold reduction in polymerization activity and increase of ethylene in copolymer.  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