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.
Model living oligomers of N-benzoyl-8-octanelactam initiated by the model octanoylium hexachloroantimonate, SbCl5, and by Ph3C+ AsF6?, were studied by 1H and 13C one- and two-dimensional nuclear magnetic resonance using the COSY, S-INEPT, COLOC, and ROESY methods. If prepared in the more solvating mixture of 1,1,2,2-tetrachloroethane (TCE) with 1,2-dichlorobenzene, the oligomers were found to be linear with the acylium ion as a living group on one end; in TCE only, the oligomers are highly branched. In both cases, the propagation center appears to be strongly coordinated to the nearest benzoylamide chain group. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
We have characterized the effective rate constants for termination/trapping (kt/t) and propagation (kp) for solvent‐free cationic photopolymerizations of phenyl glycidyl ether for conversions up to 50%. We have performed dark‐cure experiments in which active centers are produced photochemically for a specified period of time until the initiating light is shuttered off, and then the polymerization rate is monitored in the dark. This method is especially well suited for characterizing cationic polymerizations because of the long active center lifetimes. Our analysis provides profiles of the instantaneous kinetic rate constants as functions of conversion (or time). For photopolymerizations of phenyl glycidyl ether initiated with iodonium photoinitiators, kt/t and kp remain essentially invariant for conversions up to 50%. For the photoinitiator (tolycumyl) iodonium tetrakis (pentafluorophenyl) borate (IPB), the values of kt/t at 50 and 60 °C are 0.027 and 0.033 min?1, respectively. The corresponding values of kt/t for diaryliodonium hexafluoroantimonate (IHA) are 0.041 and 0.068 min?1. The values of kp at 50 °C for IPB and IHA are 0.6 and 0.4 L mol?1 s?1, respectively. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 2064–2072, 2003  相似文献   

3.
Rate constants (k1 k11, k12, k22, k21 and kt) for various steps involved in the copolymerization of propylene oxide (PO) with tetrahydrofuran (THF) have been calculated from reaction rate data obtained with the following catalyst system: (a) triphenyl-methyl cations ((C6H5)3C+) associated with hexafluorophosphate (PF6 ?), hexafluoroarsenate (AsF6 ?) and hexafluoroantimonate (SbF6 ?) gegenions; (b) antimony pentachloride (SbCl5); and, (c) boron trifluoride etherate, BF3:(C2H5)2O. The latter two systems were studied in the presence of cocatalysts. The effects of several parameters (the cocatalyst concentration and bulk size, the nature of the solvent, and the reaction temperature) on the rate constants are highlighted. The role of entropy in the initiation, propagation and termination steps is discussed in terms of solvation and desolvation processes. Based on termination activation energy considerations, the order of stability for the gegenions used in the copolymerization of PO with THF was found to be: AsF6 ? > SbF6 ? > HOBF3 ?PF6 ? > SbCl6 ?  相似文献   

4.
The electrochemical behavior of SbCl3 and SbCl5 is studied in nitromethane. SbCl3 and SbCl5 are Cl? acceptors, giving respectively SbCl4/? (log K formation=4), and SbCl6 (log K formation=15). Formal potentials of systems are determined. SbCl5 reacts with HCl, giving the solvated proton HS/+ and SbCl6/?; it is not possible to determine the formal potential of the hydrogen electrode, using HSbCl6 as an acid, because the reduction of Sbv occurs before the reduction of HS/+.  相似文献   

5.
Pulsed laser polymerization was used in conjunction with aqueous‐phase size exclusion chromatography with multi‐angle laser light scattering detection to determine the propagation rate coefficient (kp) for the water‐soluble monomer acrylamide. The influence of the monomer concentration was investigated from 0.3 to 2.8 M, and kp decreased with increasing monomer concentration. These data and data for acrylic acid in water were consistent with this decrease being caused by the depletion of the monomer concentration by dimer formation in water. Two photoinitiators, uranyl nitrate and 2,2′‐azobis(2‐amidinopropane) (V‐50), were used; kp was dependent on their concentrations. The concentration dependence of kp was ascribed to a combination of solvent effects arising from association (thermodynamic effects) and changes in the free energy of activation (effects of the solvent on the structure of the reactant and transition state). Arrhenius parameters for kp (M?1 s?1) = 107.2 exp(?13.4 kJ mol?1/RT) and kp (M?1 s?1) = 107.1 exp(?12.9 kJ mol?1/RT) were obtained for 0.002 M uranyl nitrate and V‐50, respectively, with a monomer concentration of 0.32 M. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 1357–1368, 2005  相似文献   

6.
Dilute solution viscosity measurements of nylon-6 in molten SbCl3 reveal a polyelectrolyte effect that becomes more pronounced with increased molecular weight of the polymer sample. Intrinisic viscosities result in a relationship of [η] = 2.35 × 10?6M1.45w for nylon-6 in SbCl3 at 100°C, which indicates a high polymeric chain extension in molten SbCl3 in the limit of zero concentration. Infrared (IR) and nuclear magnetic resonance (NMR) spectra indicate that a substantial fraction of the amide groups in each polymer chain remains unaffected, whereas the rest is interacted, probably, with SbCl4-ions originating from the self-ionization of SbCl3.  相似文献   

7.
Reaction Products of Chloromethoxiphosphines and Antimony (V) Chloride. Vibrational Spectra of the 1:1-adducts of Methoxiphosphoryl Compounds and Antimony (V) Chloride Chloromethoxiphosphines react with antimony(V) chloride in a redox process to yield the chloromethoxiphospllonium hexachloroantimonates(V) (CH3O)3PCl2+SbCl6? (II) and CH3OPCl3+SbCl6? (III). II, III, (CH3O)3PCl+SbCl6?(1) and (CH3O)4P+SbCl6? eliminate easily methyl chloride and give the addition compounds OP(OCH3)3·SbCl5(IV), OPCl(OCH3)2 · SbCl5 (V), OPCl2(OCH3)·SbCl5 (VI) and OPCl3·SbCl5 (VII). The vibrational spectra of IV, V nnd VI are discussed.  相似文献   

8.
Preparation of Trimercaptosulfonium Salts [S(SH)3]+AsF6? and [S(SH)3]+SbCl6? The preparation of the trimercaptosulfonium salts [S(SH)3]+AsF6? and [S(SH)3]+SbCl6? from SCl3+ salts with excessive H2S at 193 K is reported. The [S(SH)3]+SbCl6? is transferred into [S(SCl)3]+SbCl6? by reaction with Cl2 at low temperatures. The new [S(SH)3]+ cation is isoelectronic to P(PH2)3. In addition, its existence is supported by an ab-initio calculation. The results show a potential well for C3v configuration with SH bonds bended towards the top of the pyramid for the isolated ion. Also the results of a force-field calculation are reported.  相似文献   

9.
NQR Studies on Adducts of SbCl3 and SbCl5. 35Cl- and 121,123Sb-NQR Spectroscopic Studies of the Interaction between SbCl3 or SbCl5 and Phosphoryl or Thiophosphoryl Chlorides The complex formation between SbCl3 and SbCl5, respectively, and the chlorides POCl3, P2O3Cl4, and P2OS2Cl4 has been studied by NQR spectroscopy. The formation of the adducts SbCl3 · 2 POCl3 and SbCl5 · P2OS2Cl4 has been proved. Based on its 35Cl-NQR spectrum for the latter compound the constitution of a hexachloro antimonate P2OS2Cl3+SbCl6? is assumed.  相似文献   

10.
Vibrational Spectra of Trimethylphosphonium Cations (CH3)3PX+ (X = H, D) and Crystal Structures of (CH3)3PD+SbCl6? and (CH3)3PCl+SbCl6? The trimethylphosphonium salts (CH3)3PX+SbCl6? (X = H, D) and (CH3)3PH+MF6? (M = As, Sb) are prepared and characterized by vibrational and NMR spectroscopy (1H, 31P, 13C). In addition the crystal structures of (CH3)3PD+SbCl6? and (CH3)3PCl+SbCl6? are reported. (CH3)3PD+SbCl6? crystallizes in the orthorhombic space group Pnma with a = 1555(1) pm, b = 753.1(8) pm, c = 1166(1) pm Z = 4. (CH3)3PCl+SbCl6? crystallizes triclinic in the space group P1 with a = 704.6(4) pm, b = 729.5(3) pm, c = 1391.1(7) pm, α = 89.57(4)°, b? = 88.04(4)°, γ = 74.98(4)° and Z = 2.  相似文献   

11.
Preparation of Dimethyl(mercapto)sulfonium-hexachloroantimonate [(CH3)2SSH]+SbCl6? The preparation of [(CH3)2SSH]+SbCl6? from [(Ch3)2SCl]+SbCl6? and H2S at 223 K is reported. This salt is stable below 243 K and is characterized by vibrational spectroscopy.  相似文献   

12.
Summary Seven new coordination compounds are reported with the cyclicpolyether 18-crown-6 as the ligand,viz. [Mg(18-crown-6) (H2O)2](SbCl6)2, [M(18-crown-6)(MeNO6)2](SbCl6)2 with M is Ca2+ and Sr2+, [M(18-crown-6)(MeNO2)](SbCl6)2 with M is Mn2+ and Co2+, and [M(18-crown-6)](SbCl6)2 with M is Ni2+ and Zn2+.  相似文献   

13.
G. Olofsson  I. Olofsson 《Tetrahedron》1973,29(12):1711-1716
The enthalpies of interaction between H2O·SbCl5 and the oxygen bases DMA, n-Pr2O, MePrCO, and EtOAc to form ternary complexes D·H2O-SbCl5 in 1,2-dichloroethane solution have been determined calorimetrically. Reaction of water with the binary adducts D·SbCl5;, result in the formation of the same ternary complexes which has been confirmed by PMR spectroscopic experiments. The base molecules are considered to be strongly H-bonded to the H2O-SbCl5 adduct in the complexes in solution.The enthalpy of formation of H2O-SbCl5 in 1,2-dichloroethane solution has been redetermined.The enthalpies of formation of the [DH]+SbCl?6 and [D2H]+SbCl?6 complexes from the complex acid HCl-SbCl5 and Pr2O, MePrCO and (MeO)2CO have been determined and estimates of the association constant for the formation of the [DH]+SbCl?6 complexes from the D·SbCl5, adducts have been derived for D = MePrCO, EtOAc and (MeOP2CO.Measurements of the interaction between HCl-SbCl5 and the protogenic ligands PrOH and H2O were also made.  相似文献   

14.
The synthesis of a series of neodymium complexes supported on modified silica is reported. In an initial step the silanol groups were masked by a Lewis acid (BCl3, AlCl3, TiCl4, ZrCl4, SnCl4, SbCl5, HfCl4), and then a soluble arene complex Nd(η6‐C6H5Me)(AlCl4)3 formed in situ was reacted with the modified silica. The supported complexes are active and highly stereospecific for butadiene polymerization; 1,4‐cis insertion is superior by 99%. The catalyst based on a treatment of silica with BCl3 is the most efficient.  相似文献   

15.
The initiation process for the polymerization of tetrahydrofuran with (C6H5)3C+SbCl6? has been studied. Two mechanisms have been considered: a cation-addition process, and a process in which tetrahydrofuran donates a hydride ion to the cation of the initiator to form triphenylmethane. The biscarbonium salt [(C6H5)2C+C6H4CH2]2(SbCl6?)2 has been synthesized and used to initiate the polymerization of tetrahydrofuran. The results are consistent with the hydride-ion mechanism but may be inconclusive because of chain transfer. NMR experiments with 0.05–0.2M solutions of initiator in tetrahydrofuran show that triphenylmethane is rapidly produced in an amount equal to the molar amount of initiator originally present. Some NMR evidence for the presence of an acetal end group in the polymer has been obtained. It is concluded that the initiation process in this system definitely involves the formation of triphenylmethane, although a detailed, unique mechanism cannot be selected at this time.  相似文献   

16.
The cationic polymerization of ethylene oxide by trityl salts (BF4 ?, SbCl6 ?, AsF6 ?, and PF6 ? as counterions) in nitrobenzene at different temperatures has been studied. The kinetic analysis was carried out by use of an automatic manometer, and it showed that the polymerization rate constant depends neither on the counterion type nor on the initial initiator concentration. These facts allowed us to conclude that macrocations and macroion pairs have the same reactivity.  相似文献   

17.
Synthesis, Crystal Structure, and 12Sb-Mössbauer Spectrum of [SbCl2(18-Crown-6)]+SbCl6? The title compound, which has been prepared by the reaction of antimony trichloride and antimony pentachloride in the presence of 18-crown-6 in acetonitrile solution, is characterized by its 121Sb-Mössbauer spectrum and by an X-ray structure determination. Space group P21, Z = 2, 2439 observed unique reflections, R = 0.045, wR = 0.043. Lattice dimensions at ?80°C: a = 780.6(7), b = 1297.5(9), c = 1 278.5(10) pm, β = 100.56(7)°. The structure of [SbCl2(18-crown-6)]+SbCl6? contains cations in which the antimony atom in the first coordination sphere is surrounded in a ?-trigonal-bipyramidal fashion by two oxygen atoms of the crown ether in axial position as well as in the equatorial position by the two chlorine atoms and the lone electron pair.  相似文献   

18.
The kinetics of polymerization of 1, 3-dioxolane (DiOX) initiated by (C2H5)3O+SbCl6 and SbCl5 has been studied and the elementary stages of the process have been considered. The polymerization of DiOX by (C2H5)3O+SbCl6-is shown to proceed at a steady rate to high conversion. A constant concentration of active centers in the system is maintained due to the equal rates of decomposition of active centers and disproportionation. The nonsteady-state character of DiOX polymerization initiated by SbCl5is associated with a relatively lower stability of the counter-ion SbCl5 OR? compared with SbCl6. The initiation of DiOX polymerization by (C2H5)3O+SbCl6 proceeds without hydride-transfer reactions, and the concentration of active centers in the system is determined not by processes taking place in the initiation stage, but by the existence of a definite kind of equilibrium with the participation of active centers.  相似文献   

19.
For a temperature range of −11.8–92.6°C, the propagation rate constant kp of styrene has been determined with the use of pulsed-laser polymerization (PLP). The temperature dependency of the obtained kp data was evaluated using the Arrhenius equation. The NLLS error-in-variables method (EVM) is recommended for this fit. The resulting activation energy is 32.6 kJ mol−1 and the pre-exponential factor is 107.66 dm3 mol−1 s−1. A joint confidence interval for these parameters is given. © 1996 John Wiley & Sons, Inc.  相似文献   

20.
Polyacetylene, (CH)x, has been doped with trimethyloxonium hexachloroantimonate, (CH3)3O+SbCl(1), in dichloromethane and acetonitrile. The maximally doped (CH)x films have moderate conductivities [σRT(CH2Cl2) = 10, σRT(CH3CN) = 0.7 Ω?1 cm?1]. Reactions between 1 and (CH)x CH2Cl2 or CH3CN were followed in situ by 1H nuclear magnetic resonance spectroscopy and x-band electron spin resonance spectroscopy. It was found that the reactions in the two solvents are different. In dichloromethane the dopant is SbCl5, which forms from the decomposition of 1, and doping proceeds by electron removal from (CH)x chains. Based on the ESR signal loss, an estimate can be made of the diffusion rate of SbCl5, into the (CH)x fibrils in CH2Cl2; it is found to be ca. 10?17 cm2/s. In acetonitrile the dopant appears to be either CH3CNCH, H+, CH, or a combination of one or more of these dopants. It is postulated that the CH3CNCH, CH, and/or H+ dopant covalently binds to the (CH)x chain. X-ray photoelectron spectra show that films doped with excess 1 in both solvents have approximately one SbCl per 33 CH units.  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