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.
By means of 19F-NMR spectroscopy it was possible to detect and estimate the hexafluorophosphate ion (PF) in the THF polymerization, initiated by PF5 in methylene dichloride at 25°C. The observed variations with time of [PFPF], [PF5·THF], and [PPF] were analysed kinetically, and by means of the resulting equations, the rate constants for initiation, propagation and termination were evaluated. Comparison of our kp value with those of others indicates that our method is sound, and we conclude that it should be applicable to other similar systems.  相似文献   

2.
The extent of reaction of functional groups (pA), the extent of conversion of monomer molecules (qM), the true relative polymer yield (yp), the constitutional repeating unit-based apparent relative polymer yield (y), and the monomer-based apparent relative polymer yield (y) are defined and their interrelationships given. The use of y instead of yp overestimates the true yield, except in the limit of high degrees of polymerization. Use of y, on the other hand, always underestimates the true yield, especially at infinite degrees of polymerization. The correct expression for the calculation of the mass fraction of i-mers in reactants with a Schulz-Flory “most probable” distribution is given and shown to become identical with the usual approximation in the limit of vanishingly small mass of the leaving molecule.  相似文献   

3.
The mechanism of the cationic polymerization of cyclic sulfides, initiated with triethyloxonium salts is reported. Differences due to changing the counter-ion from BF to SbCl are discussed. Initiation consists of the alkylation of monomer forming a cyclic sulfonium ion which is the active species in the propagating step. A generally occurring termination reaction is the formation of non-strained (linear or cyclic) sulfonium ions. In the case of epithiopropane (propylene sulfide) these can slowly re-initiate the polymerization by an intramolecular reaction forming the 3-membered cyclic sulfonium salt. This re-initiation is not possible with the thietanes because of the greater difficulty to form 4-membered cyclic sulfonium salts. When SbCl is the counter-ion, an additional, more drastic termination can occur most probably by reaction of the growing chain with the counter-ion, thereby forming an alkyl chloride and SbCl5. After the polymerization of propylene sulfide with BF as the counter-ion, the polymer degrades to low molecular weight oligomers, predominantly the cyclic tetramer. This degradation is a back-biting reaction occuring via sulfonium salts. With SbCl as the anion, the sulfonium salts are destroyed by the termination reaction and degradation does not occur. It was possible to follow the concentration of the growing species (4-membered cyclic sulfonium salt) during the polymerization of 3,3-dimethylthietane by means of 300 MHz NMR spectroscopy. By measuring the rate constants of propagation for different initiator concentrations or in the presence of different amounts of an indifferent electrolyte, it was possible to calculate separate rate constants for propagation via free ions (k) and via ion-pairs (k). The ratio k/k was about 60 when BF was the anion and 35 when it was SbCl.  相似文献   

4.
Following our earlier work on the polymerization of lactones involving crowned cations, kinetics of the anionic polymerization of ?-caprolactone (?CL) with K+ · (dibenzo-18-crown-6 ether) (K+DB18C6) counterion was studied calorimetrically in THF solution in the temperature range from 0 to 20°C. Dissociation constants of CH3(CH2)5O?K+DB18C6, modelling the active centers, were determined conductometrically: KD (20°C) = 7,7 · 10?5 mol · dm?3, ΔH = 9,3 ± 0,2 kJ · mol?1, ΔS = ?47 ± 2J · mol?1 · K?1. From kinetic measurements and from measurements of the dissociation constant of CH3(CH2)5O? K+DB18C6, rate constants of propagation via macroions and via macroion pairs were determined. Activation parameters for propagation via these species are equal to: ΔH = 39,2 ± 0,2 kJ · mol?1, ΔS = ?63 ± 1 J · mol?1 · K?1, ΔH = 13,7 ± 0,1 kJ · mol?1, ΔS = ?185 ± 2 J · mol?1 · K?1. At 20°C, k = 3,50 · 102 dm3 · mol?1 · s?1 and k = 5,2 dm3 · mol?1 · s?1. Due to the large difference of ΔH for propagation via macroions and macroion pairs (vide supra), the isokinetic point (k = k) would appear at ?65°C.  相似文献   

5.
The possibility of determination of the propagation and termination rate constants kp and kt, resp., as well as their ratios kp/kt and kp/k in homogeneous radical polymerization is shown using the capacity flow method. A theoretical analysis is carried out and relatively simple equations are introduced. The essential point is that the life time of the growing polymer chain is obtained graphically from the residence time of the reactants in the reactor vessel, which is a given quantity. As an experimental example the polymerization of methyl methacrylate initiated by benzoyl peroxide in benzene is investigated in a flow reactor with perfect mixing at 80°C. It is characteristic that the process can be followed by means usually applied for studying slow reactions. The degree of conversion is measured turbidimetrically and gravimetrically, whereas the initiation rate is analysed iodometrically. Thus, through a numerical linear approximation by the method of least squares kp/kt = (2,28±0,45)·10?5, kp/k = (1,50±0,22). 10?1 are found from the experimental data, and hence kp = (9,95±0,83)·102 l mol?1 s?1 and kt = (4,36±0,49)·107 l mol?1 s?1 are obtained.  相似文献   

6.
An attempt has been made to clarify the effect of molecular weight distribution (MWD) in the original polymer on the separation characteristics in successive solutional fractionation (SSF) and comparisons were made with successive precipitational fractionation (SPF). Polymers with Schulz-Zimm distribution and Wesslau distribution of the degree of polymerization (ratio of the weight- to number-average degree of polymerization, X?/X? = 2 to 5) were brought into solution, and cooled to cause precipitation according to the modified simulative procedure of Kamide and Sugamiya, based on the Flory-Huggins's theory of dilute solutions of polymers. No double-peak distributions were observed in the fractions obtained by SSF under the conditions which give double-peak distributions in the case of SPF. The minimum value of X?w in the fractions obtained in a given run is almost the same for SSF and SPF, the minimum increasing with decreasing X?/X? (the suffix zero denoting the original polymer). The ratio X?w/X?n in the fractions is not greatly influenced by X?/X? except for the first few fractions. The standard deviation σ′ of the MWD in the fractions decreases with increasing X?/X? in an initial stage, while this situation is reversed in the later stages. In general SSF furnishes sharp fractions regardless of X?/X?.  相似文献   

7.
Kinetic studies of the anionic polymerization of 2-diethylamino-1,3,2-dioxaphosphorinane were performed in THF solution with (CH3)3SiO?K+ as initiator at temperatures close to r.t. Initiation involves nucleophilic attack of the anion on P atom in the monomer molecule. Breaking of the P? O bond leads to an alcoholate anion as the growing species. Polymerization was shown to proceed via macroion-pairs and to be nearly living; e.g. at r.t. for every 250 propagations there is one termination. Rate constant of propagation k = 3,4 ± 0,31·mol?1·s?1 at 25°C, ΔH = 13,3 kcal·mol?1 and ΔS = ?32,2 cal·mol?1·K?1. The ratio k/k was determined by solving a kinetic scheme involving propagation and termination. It was shown that termination consists in the alcoholate anion attack on P in either polymer or monomer molecule with expulsion of (C2H5)2N? anion and formation of a P? O bond. The dialkylamide anions cannot reinitiate polymerization. In solving the kinetic scheme it was assumed that termination involving both polymer and monomer proceeds with rate constants equal to each other.  相似文献   

8.
The termination mechanism in the polymerization of a series of analogous di(meth)-acrylates differing only by the type of the heteroatom (S or O) in the ester group was investigated. The experimental method was based on differential scanning calorimetry measurements of photochemically initiated polymerization. The ratio of the bimolecular termination and propagation rate constants k/kp and/or the monomolecular termination rate constant k were determined according to three termination models (monomolecular, bimolecular, and mixed). Statistical techniques were used for choosing between the models to find out by which mechanism (monomolecular, bimolecular or mixed) the termination occurs. The results obtained prove than monomolecular reaction is an important termination way in the polymerization of multifunctional monomers and must be taken into account when kinetic constants are determined, even at low degree of conversion. The sulfur-containing monomers are characterized by much lower k values than the analogous sulfur-free monomers.  相似文献   

9.
An attempt has been made to account for the occurrence of reverse-order fractionation (ROF) phenomena in successive precipitational fractionation (SPF) in terms of the newly established rigorous fractionation theory, assuming complete thermodynamic equilibrium conditions between a polymer-rich phase and a polymer-lean phase and to throw light on operating conditions under which ROF occurs. For this purpose, the simulation technique was employed. ROF occurs not only between 1st and 2nd fractionation steps but also between two successive higher order steps even under thermodynamic equilibrium state. The molecular weight distribution of the original polymer contributes very sensitively to ROF: For the Schulz-Zimm type polymer ROF is highly observed at a large amount of fraction Q in a narrow range of initial concentration v whereas for the Wesslau type polymer large ROF occurs at small Q over a relatively wide v range. In general, there exist appropriate ranges of the weight-average degree of polymerization of the initial polymer X? the solvent nature expressed by the concentration dependence of the polymer/solvent interaction parameter p, and the initial polymer concentration v for ROF. A broad original polymer gives rise to large ROF. The theoretical predictions obtained here can explain very clearly the experimental results by Fujisaki and Kobayashi concerning the effect of X?, p, and v on ROF in SPF of polyacrylonitrile.  相似文献   

10.
The kinetics of the anionic polymerization of styrene were investigated under pressure (1≤p/bar < 1800) with Na+ as counter ion in tetrahydropyran (THP) as solvent and with Cs+ as counter ion in 1,2-dimethoxyethane (DME) as solvent. The results yielded the activation volume of the contact ion pair ΔV and the sum (ΔV + ΔVcs) of the activation volume of the solvent separated ion pair ΔV and the volume change upon formation of solvent separated ion pairs from contact ion pairs ΔVcs. The numerical values are negative. The activation volume of the solvent separated ion pairs could be estimated.  相似文献   

11.
A compensation effect exists between the quantities (ΔH ? ΔH) and (ΔS ? ΔS) in the free radical polymerization of a monomer in different solvents ΔH, ΔH, ΔS, and ΔS are the activation enthalpies and entropies, resp. for the formation of isotactic and syndiotactic dyads. The quantities ΔΔH and T0 are by definition independent of the temperature of polymerization and other polymerization conditions and thus a pair of constants characteristic for each monomer. A linear relationship between ΔΔH and T0 has been found for acrylic and vinyl monomers each. Both true activation and conformational effects seem to be responsible for the stereocontrol in free radical polymerizations.  相似文献   

12.
The initial rate of the polymerization of methyl acrylate photoinitiated by 4-dimethylamino-4′-isopropylbenzophenone ( 2 ) varies linearly with monomer concentration and the square-root of light intensity. For small absorbances, the initial polymerization rate, Rp, varies linearly with the square-root of the photoinitiator concentration. For high absorbance values, Rp is independent of the photoinitiator concentration in a well-stirred system and decreases as the photoinitiator concentration increases in an unstirred system. This marked differences between the polymerization kinetics in well-stirred and unstirred systems is discussed. The ratio of the propagation rate constant and the square-root of the termination rate constant, kp/k, and the quantum yield of photoinitiation were determined and related to the results previously obtained in a study on the photoreactivity of 2 .  相似文献   

13.
Anionic polymerizations of butyl cyanoacrylate were initiated in tetrahydrofuran (plus a few experiments in 1,2-dimethoxyethane) by the salts tetrabutylammonium hydroxide, bromide, acetate and three substituted acetates. The hydroxide gives near-ideal ‘living polymerization’ kinetics, with kp close to 106l·mol?1. s?1 at 20°C. The kinetics of the reactions initiated by the acetates and bromide are analysed by the slow-initiation-no-termination theory, using values of the initiation rate constants evaluated in Part 1. The kp values derived are in the same range as those from the OH-initiated reactions and those of the zwitterionic polymerizations initiated with covalent bases, i.e., tertiary phosphines and amines. A ca. 4-fold variation of kp with concentration of active species is given a speculative analysis in terms of dissociation from paired to free ions, yielding tentative estimates for k ≈ 105 and k ≈ 107l·mol?1·s?1 in THF at 20°C. Molecular weights were all high M n ≈ 106, with M w/M n ≈ 2.  相似文献   

14.
1,1,4,4-Tetraphenyl-2-tetrazene ( 1 ) inhibits the radical polymerization of styrene initiated by 2,2′-azoisobutyronitrile. The inhibition period is proportional to the concentration of 1 . The rate of the initiation of polymerization and the ratio of the rate constants of propagation and termination reactions k/2kt, were calculated from the consumption rate of the inhibitor. Some aromatic amines did not effect the inhibition period and the polymerization rate. Tetraphenylhydrazine showed a retarding effect.  相似文献   

15.
Four topics related to the cationic polymerization of cyclic acetals and mainly based on the author's results are reviewed.
  • (1) The kinetics of the H? transfer from linear and cyclic acetals and ethers to the triphenylmethyl cation were examined. A linear correlation between the 1H-NMR shift of the H-atom to be transferred as H? anion and the enthalpy of activation (ΔH) was observed for H? donors which are linear and cyclic-planar or deviate slightly from planarity. Puckered rings, like 1,3,5-trioxane were found to be very poor H? donors. A linear relationship between ΔH and ΔS was observed for all studied compounds.
  • (2) The equilibria carbenium ion ? oxonium ion involving the unpaired triphenylmethyl cation and different linear and cyclic acetals and ethers were studied. A linear relationship between the basicity of the involved nucleophile (expressed as pKa or pKb) and the enthalpy of equilibrium (ΔH) was found.
  • (3) The chemistry of the elementary reactions leading eventually to the formation of the active species in the polymerization of 1,3-dioxolane was investigated. The latter was used as a model monomer and initiated with triphenylmethyl salts with various stable anions like AsF and SbF or with the unstable SbCl. The importance of the ambident reactivity of 1,3-dioxolan-2-ylium cations was pointed out and the kinetics of cationation via ring opening of these salts was described. 1,3-Dioxolan-2-ylium salts were shown to be sufficiently stable to be slowly transformed into the linear cations (k?cat ≈ 10?3 to 10?4 mol?1 dm3 s?1 at 25°C). This slow reaction was partially responsible for the induction periods observed in the polymerization of 1,3-dioxolane.
  • (4) The kinetics and mechanism of the chain propagation of 1,3-dioxolane were described and a structure ( 20 ) was proposed for the growing centres. It was indicated that the growing species, being the tertiary oxonium ions use the polymer segments as ligands and in contrast to the oxycarbenium ions do not abstract hydride anions (H?).
  相似文献   

16.
The influence of the temperature of the melt T1 on the kinetics and the morphology of a semicrystalline polymer (poly(oxymethylene)) was investigated using thermal analysis and optical microscopy. The thermodynamic melting point T and the enthalpy of melting at thermodynamical equilibrium ΔH were determined by extrapolation of the graphs Tf = f(Tc) and ΔHf = f(Tc); (T = 198°C, ΔH = 251 J/g). For different temperatures of the melt (T1 = 185°C, 195°C, 205°C), isothermal and non-isothermal crystallizations were analysed using the Avrami and Ozawa equations. Nucleation and spherulitic growth in this polymer were studied by using optical microscopy at elevated temperatures. Using different analyses, we observed initial nucleation followed by spherulite growth with the following influence of the temperature of the melt on the distribution and the number of spherulites: T1 < T produces many small spherulites; T1 > T gives rise to few large spherulites.  相似文献   

17.
Salts of stable carbocations such as triphenylmethyl and tropylium ions were used conveniently to estimate free ion propagation rate coefficients (k) for cationic homopolymerization of a series of alkyl vinyl ethers and N-vinylcarbazole. Methyl vinyl ether and 2-chloroethyl vinyl ether gave values of k approximately one order of magnitude less than those of branched alkyl derivatives, and an explanation in terms of conformational effects is presented. Results are compared with those from radiation induced polymerizations, and with data for related hydride ion abstraction processes, in an attempt to evaluate solvent effects on free ion reactivity. The scope and applicability of stable carbocations to cationic polymerizations are surveyed and complicating side reactions exemplified by the systems: xanthylium ion—1,1-diphenylethylene, tropylium ion—indene, xanthylium ion—THF, C6H5CO+SbCl—THF, and by the effect of perchlorate anions on chain transfer in polymerization of N-vinylcarbazole.  相似文献   

18.
Experimental details are given of attempts to enumerate the binary ionogenic equilibria (B.I.E.) of 1-chloro-1-methylethylbenzene ( 1 )/BCl3, 1,4-bis(1-chloro-1-methylethyl)benzene ( 2 )/BCl3 and 1,3,5-tris(1-chloro-1-methylethyl)benzene ( 3 )/BCl3 in CH2Cl2. Due to chemical reaction (dimerisation or polymerisation) no experimental values for the B.I.E. constants could be obtained. A Born-Haber cycle is constructed to estimate the relative sequence of the overall B.I.E. constants. A similar treatment for 2-chloro-2methylpropane as a thermodynamic model for α,ω-dichloropoly(2-methylpropene) ( 4 ) suggests that the overall B.I.E. constant for these polymers is somewhat smaller than those for 1 and 2 but greater than that for 3 . Using 2 /BCl3 as initiator for the polymerisation of 2-methylpropene (IB) it is shown, that the degree of polymerisation of 4 can be controlled within the limits 10 < DP < 100. It is shown that 4 can also act as an initiator for the polymerisation of IB, that these polymerisations involve only free ion propagation and, from a kinetic analysis of these polymerisations, that: (k)2/k = 12 1 · mol?1 · s?1, k = 1,2 · 10?3 l · mol?1 · s?1, k [P] = 1,7 · 10?3 s?1, and k/(k K) = 102. The same analysis demonstrates that the self-ionisation of BCl3 can be neglected in terms of any influence on the molar mass of the products. Experiments are also described which show that 2-chloro-2-methylpropane is not suitable as a substitute initiator for IB, but that 2-chloro-2,4,4-trimethylpentane is a useful model for 4 as an initiator for the polymerisation of IB.  相似文献   

19.
The kinetics of hydroxymethylation of melamine (2,4,6-triamino-1,3,5-triazine) with formaldehyde to form N-(hydroxymethyl)melamine (2,4-diamino-6-hydroxymethylamino-1,3,5-triazine) was investigated in aqueous hydrogen phosphate/phosphate buffer solutions in the range of pH 10,51 to 12,20 at 20°C, determining unreacted melamine by means of HPLC. As a result, this reaction is subject to a general base catalysis, and the second order rate constant k is expressed by k = k′ + kA ? [A?] + k[A?]2/[HA] (or k [A?][OH?]), where HA and A? denote acid and basic constituents of the buffer, and k′ is the rate constant in unbuffered media. This rate equation is compared with those reported previously.  相似文献   

20.
In contrast to the usual method of determination of the principle ratio in radical polymerization, k/kTe, which is based on the comparison of the rate of polymerization with the concentration of initiator, a method was used consisting in the representation of loge([M]0/[M]) versus (1 ? e?kdt/2). The results obtained with styrene, vinyl acetate, acrylic acid, methacrylic acid, and 3,3,4,4,5,5,6,6,7,7,8,8,8-tridecafluorooctyl acrylate ( 1 ) were found to be in good agreement with those already obtained by the other method.  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