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 78 毫秒
1.
The reaction of some diaryl-substituted olefins [1,1-diphenylethylene (DPE), 1,1,3,3-tetraphenyl 1-butene (TPB) and 3-phenylindene (PI)] with trifluoromethanesulphonic acid has been studied in dichloromethane by means of u.v. and 1H NMR spectroscopies. All carbocations are stable at temperatures below -30°. In the case of DPE at ?70°, the carbocation initially formed is dimeric but is slowly transformed into a monomeric cation. The initiation step is too fast to be studied by conventional means except for TPB where a first order rate with respect to the olefin has been observed.For DPE and TPB, when the ratio olefin/acid is high, a 33% yield in carbocations with respect to CF3SO3H is obtained for the whole range of acid concentration investigated (10?4 ?1 mol 1?1). In the case of PI, the yield is 50%. 1H NMR spectra clearly show that some acid remains unreacted in the presence of olefin. The large downfield shift of unconsumed acid is consistent with the existence of complex anions CF3SO?3, (HOSO2CF3)2 and CF3SO?3, HOSO2CF3 resulting from strong hydrogen bonds between the triflate anion and its conjugate acid.  相似文献   

2.
The initiation reaction of 1,1-diphenylethylene (DPE) dimerization by TiCl4 in dichloromethane has been studied at ?30 and ?70°C in the presence and absence of added cocatalyst. When water was added to the medium, the concentration of active centers was proportional to cocatalyst concentration. When carbocations R+ are formed without a cocatalyst, their usually low concentration depends on initial concentrations of reagents; thus |R+| ∝ |C|0/|M|0, which indicates that the olefin has an inhibiting effect. The simplest hypothesis applicable to these two observations is the existence of a secondary reaction parallel to initiation which traps the initiator in an inactive form; for example, as a complex with the monomer. The two kinetic schemes, which are proposed according to the stoichiometry of the inactive complex (bimolecular C,M or trimolecular C,M2), support mechanisms that would involve solvent cocatalysis, electronic transfer, or direct addition of TiCl4 on the monomer double bond. Consideration of these rival theories reveals that the most likely is the direct addition of TiCl4 on monomer, but unfortunately the known instability of the corresponding organotitanium intermediates does not allow their characterization.  相似文献   

3.
The chemical microstructure of arylene–ether–ketone copolymers of terephthaloyl chloride (TCl), 1,4-diphenoxybenzene (DPB), and diphenyl ether (DPE) has been characterized by 13C-NMR. Copolymers synthesized via two Friedel–Crafts reaction have been investigated; the first reaction uses HF and BF3 to catalyze polymerization, and the second uses LiCl to moderate the activity of the Lewis-acid catalyst, AlCl3. The HF/BF3 approach results in random copolymers, in which the TCI displays no preference in reacting with either DPB or DPE. The buffered AlCl3 approach yields somewhat blocky copolymers, in which the DPB and DPE tend to segregate. The degree of segregation in these materials has been quantified by formulas for the number-average block lengths.  相似文献   

4.
E. M. F. of the Cell, Cd-Hg (2-phase)/CdAc2(m), Hg2Ac2(s)/Hg was measured at 20°, 25°, 30° and 40°C. The standard e. m. f. of the cell, Cd/CdAc3(m), Hg2Ac2(c)/Hg was evaluated as E°=1.1500?11.09×10?4T+1.06×10?8T2 The thermodynamic data of the reaction, Cd(c) + Hg2Ac2(c)=2Hg(l)+Cd++(aq)+2Ac?(aq) at 25°C were estimated as ΔF°=?42,139, ΔH°=?48,698 cal mole?1 and ΔS°=?22.0 cal deg?1 mole?1 at 25°C. The thermodynamic data for the formation of Hg2Ac2(s) were evaluated as ΔFf°=?202.3, ΔHf°=?154.5 Kcal mole?1 and S°=72.9 cal deg?1 mole?1. From measurements of the heats of solution of CdAc2·2H2O in aqueous solution, the relative partial molal enthalpies of cadmium acetate in aqueous solution were estimated.  相似文献   

5.
Abstract

The reaction between tetraphenylphosphonium chloride and hydroxide or deuteroxide anions was studied kinetically in a series of dimethylsulphoxide-water mixtures at several temperatures. The rate is first-order in the phosphonium cation and second-order in the hydroxide or deuteroxide anions. The reaction shows a dramatic increase in rate, up to about 1010 times, as the DMSO content is increased. The rate enhancement is attributed to a considerable drop in activation energy affected not only through an increased desolvation of reactant anions, but also through an increase in solvation of the transition state, brought about by gradual addition of DMSO. The kinetic solvent deuterium isotope effect in 60% DMSO-40% D2O is strongly dependent on temperature. The rate constant in the latter solvent mixture is represented by k i = 11.9 e ?12700/RT l 2 mole?2 sec?1 as compared to k i = 19.0 e ?22500/RT l 2 mole?2 sec?1 in the corresponding 60% DMSO-H2O mixture. The thermodynamic parameters of activation show strong dependence on solvent composition and are related to structural changes and solvation power of the reaction medium.  相似文献   

6.
The controlled cationic polymerization of styrene using CumOH/AlCl3OBu2/Py initiating system in a mixture CH2Cl2/n‐hexane 60/40 v/v at ?40 and ?60 °C is reported. The number‐average molecular weights of the obtained polystyrenes increased with increasing monomer conversion (up to Mn = 85,000 g mol?1) although experimental values of Mn were higher than the theoretical ones at the beginning of the reaction that was ascribed to slow exchange between reversible‐terminated and propagating species. The molecular weight distribution became narrower through the reaction and leveled of at the value of Mw/Mn = 1.8–2.0. A kinetic investigation revealed that the rate of polymerization was first‐order in AlCl3OBu2 concentration meaning that monomeric counteranion (AlCl3OH? or AlCl) involved in the initiation and propagation steps of the reaction. It was also found that the rate of polymerization decreased with lowering temperature, which could be attributed to a decrease in concentration of free Lewis acid (AlCl3), the true coinitiator of polymerization, because of an increase in the tightness of its complex with dibutyl ether. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 3736–3743, 2010  相似文献   

7.
The study of D(?)-ribose complexing with calcium in aqueous solutions less than 1.64 × 10?1M by potentiometric measurements with a calcium selective electrode afforded the value of K1 = 1.70 liters × mole?1 (SD = 1.05 × 10?3). Numerical analysis indicated that complex species with 1:1 and 1:2 calcium to D(-)-ribose ratios are present simultaneously: k1 = 1.13 liters × mole?1 and K2 = 8.47 liters × mole?1 (SD = 0.95 × 10?3).In methanolic medium 1.24 × 10?2M with regard to calcium chloride both stoichiometric proportions were evidenced. A large error accompanying the stability constant K1 = 28 kg × mole?1 (RSD = 82%) renders unreasonable the K2 value obtained from the product K1 × K2 = 96.5 kg2 × mole?2.The results are discussed with respect to the data published for more concentrated (1.27 M) aqueous solutions obtained on the basis of 1H-NMR spectroscopic investigations.  相似文献   

8.
Abstract

Infrared spectra (4000–200 cm?1) have been reported for Ni(DPE)X2 where X is Cl, Br and I and DPE is 1,2-bis(diphenylphosphino)ethane. The Ni[sbnd]X and Ni[sbnd]P stretching bands have been assigned based on the observed isotopic shifts due to the 58Ni-62Ni substitution. The Ni[sbnd]X stretching frequencies are always lower and the Ni[sbnd]P stretching frequencies are always higher in the cis-complexes such as Ni(DPE)X2 than in the corresponding trans-complexes such as Ni(PEt3)2X2. These differences between cis and trans configurations have been attributed to the strong trans-effect of phosphine ligands.  相似文献   

9.
Polymerization of acrylonitrile photoinitiated by naphthalene, anthracene, phenanthrene, and pyrene is accelerated by an admixture of zinc (II) chloride, acetate, or nitrate. The effect of zinc (II) salts on the rate of pyrene-photoinitiated polymerization of acrylonitrile leads to an increase in this rate in the order Zn/OCOCH3/2 < ZnCl2 < Zn/NO3/2. The maximum polymerization rate is achieved at the molar ratio [ZnCl2]/([ZnCl2] + [pyrene]) approximately 0.7. In contrast to the photoinitiated polymerization of acrylonitrile, the methyl methacrylate admixture of zinc (II) chloride exerts a smaller effect on the polymerization rate. In the pyrene-photoinitiated polymerization of styrene an admixture of zinc (II) chloride retards the polymerization rate. Fluorescence of aromatic hydrocarbon in the system acrylonitrile–aromatic hydrocarbon is efficiently quenched by zinc (II) chloride. Stern–Volmer constants determined for pyrene (80 dm3 mole?1), phenanthrene (66 dm3 mole?1), and naphthalene (49 dm3 mole?1) are higher by about 2–3 orders of the Stern–Volmer constants for fluorescence quenching of aromatic hydrocarbons by acrylonitrile in the absence of ZnCl2. The fluorescence of anthracene in acrylonitrile is not quenched by ZnCl2. The acceleration effect of Zn (II) salts on the polymerization of acrylonitrile photoinitiated by aromatic hydrocarbons depends on two factors: an increase in the ratio of the rate constant of the growth and termination reactions, kp/kt, and an increase in the quenching constant of fluorescence of aromatic hydrocarbon, kq, by the complex {acrylonitrile…ZnCl2}. ZnCl2 thus influences both the growth and initiation reactions of the polymerization process.  相似文献   

10.
The reaction of cyclohexyl isocyanate with phenylglycidyl ether was selected as model reaction for the synthesis of cycloaliphatic isocyanate-based poly(2-oxazolidone)s. The selectivity of AlCl3 and AlCl3-triphenylphosphine oxide (AlCl3–TPPO) and AlCl3-hexamethylphosphoric triamide (AlCl3–HMPA) complexes was studied for 2-oxazolidone formation. The reaction products were identified by means of the melting point, 1H-NMR, and IR spectroscopy. The kinetics of the model reaction was studied using AlCl3-TPPO in o-dichlorobenzene at 120 and 140°C.  相似文献   

11.
The gas phase, nitric oxide catalyzed positional isomerization of 3-methylene-1,5,5-trimethylcyclohexene (MTC) into 1,3,5,5-tetramethyl-1,3-cyclohexadiene (TECD) has been studied for temperatures ranging between 296° and 425°C. The major reaction was first order with respect to nitric oxide and to MTC. The major side product, mesitylene, usually amounted to less than 10% of the TECD isomer formed. Only at high temperatures and large conversions has up to 20% been observed. Conditioned pyrex or quartz vessels coated with KCl have been used. The nitric oxide catalyzed isomerization is apparently a homogeneous process, as demonstrated by the insensitivity of the observed rate constants towards a 15-fold increase in the surface to volume ratio of the reaction vessels. However, a residual, presumably heterogeneous, thermal isomerization of the starting material could not be eliminated. Good mass balances were obtained for both NO and hydrocarbons. After correcting for the thermally induced conversion the observed rate constants for the nitric oxide catalyzed isomerization yield log k1 (1 mole?1 sec?1) = (10.7 ± 0.2) – (37.3 ± 0.9)/θ where θ is 2.303 × 10?3 RT (kcal mole?1). Plotting log k1 versus the ratio of the starting materials (MTC/NO)0 it was found that for temperatures ≥ 365°C the rate constants were systematically too high. Using extrapolated values for the higher temperature range yields the more reliable corrected Arrhenius equation log k = 8.6 – 31.7/θ. The reaction mechanism is outlined and the implications with respect to the stabilization energy generated in the MTC? radical intermediate and the activation energy of the backreaction MTC? + HNO are discussed. Using for the activation energy E?1 of the backreaction (R? + HNO) a literature value of 9.2 ± 0.9 kcal mole?1 reported for the cyclohexadiene? 1,3? system, this yields 23.4 ± 2 kcal mole?1 for the stabilization energy in the methylenecyclohexenyl radical, which is to be compared with the corresponding values for the allyl (10.2 ± 1.4), methallyl (12.6 ± 1) pentadienyl (15.4 ± 1) and cyclohexadienyl (24.6 ± 0.7) radicals. The pre-exponential factor agrees well with the value of (8.4 ± 0.2) reported by Shaw and co-workers for the similar reaction of NO with 1,3-cyclohexadiene. It is noteworthy that HNO, acting as sole hydrogen donor in the system, is surprisingly stable under the reaction conditions used. Nitrous oxide, HCN, H2O and N2 are observed in the product mixture of experiments carried out to high conversions at higher temperatures.  相似文献   

12.
The cationic polymerization of isobutylene using 2‐phenyl‐2‐propanol (CumOH)/AlCl3OBu2 and H2O/AlCl3OBu2 initiating systems in nonpolar solvents (toluene, n‐hexane) at elevated temperatures (?20 to 30 °C) is reported. With CumOH/AlCl3OBu2 initiating system, the reaction proceeded by controlled initiation via CumOH, followed by β‐H abstraction and then irreversible termination, thus, affording polymers (Mn = 1000–2000 g mol?1) with high content of vinylidene end groups (85–91%), although the monomer conversion was low (≤35%) and polymers exhibited relatively broad molecular weight distribution (MWD; Mw/Mn = 2.3–3.5). H2O/AlCl3OBu2 initiating system induced chain‐transfer dominated cationic polymerization of isobutylene via a selective β‐H abstraction by free base (Bu2O). Under these conditions, polymers with very high content of desired exo‐olefin terminal groups (89–94%) in high yield (>85%) were obtained in 10 min. It was shown that the molecular weight of polyisobutylenes obtained with H2O/AlCl3OBu2 initiating system could be easily controlled in a range 1000–10,000 g mol?1 by changing the reaction temperature from ?40 to 30 °C. The MWD was rather broad (Mw/Mn = 2.5–3.5) at low reaction temperatures (from ?40 to 10 °C), but became narrower (Mw/Mn ≤ 2.1) at temperatures higher than 10 °C. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

13.
Study of n-butane pyrolysis at high temperature in a flow system allows measurement of the sum of the rate constants of the initiation reactions and of the Arrhenius parameters of the reactions Established data for k1/k2 allow estimation of k1 for 951°K and this, with recent thermochemical data, yields the result log k?1 (l.mole s?1) = 8.5, in remarkable agreement with a recent measurement [20] but over si×ty times smaller than conventional assumption. The product k3k4 (l.2mole?2s?2) is found to be associated with the Arrhenius parameters log (A3A4) = 21.90 ± 0.6 and (E3 + E4) = 38.3 ± 2.7 kcal/mole. These values are much higher than would be e×pected on the basis of low temperature estimates. Independent evaluation gives log A4 = 10.5 ± 0.4 (l.mole?1s?1) and E4 = 20.1 ± 1.7 kcal/mole, hence log A3 = 11.4 ± 0.8 (l.mole?1s?1) and E3 = 18.2 ± 3.2 kcal/mole. These values are shown to be entirely consistent with a wide range of results from pyrolytic studies, and it is argued that they further confirm the view that Arrhenius plots for alkyl radical–alkane metathetical reactions are strongly curved, in part due to tunneling and, appreciably, to other as yet unidentified effects. Since there is published evidence that metathetical reactions involving hydrogen atoms show even greater curvature, it is suggested that this may be a characteristic of many metathetical reactions.  相似文献   

14.
The reaction of AlCl3 with Li2PR (R = SiiPr3, SiMeiPr2) in a mixture of heptane and ether yields in the polycyclic compounds [(AlCl)43‐PR)2(μ‐PR)2(Et2O)2]( 1a : R = SiiPr3; 1b : SiMeiPr2) with a ladder‐shaped Al4P4 core. The coordination sphere of the outer aluminium atoms in these compounds is completed by ether ligands. In contrast, the reaction of AlCl3 with Li2PSiiPr3 in pure heptane yields in the formation of the hexagonal prismatic compound [(AlCl)63‐PSiiPr3)6]( 2 ). 1 and 2 were characterized by single crystal X‐ray diffraction analysis as well as by 31P{1H} and 27Al NMR spectroscopy. The structure determining effect of the solvent can be rationalized by quantumchemical calculations, which also show that the hexagonal prismatic structure is the most stable of the investigated oligomers in absence of ether.  相似文献   

15.
In the gas phase, cis,trans-1,5-cyclooctadiene (\documentclass{article}\pagestyle{empty}\begin{document}$ {\mathop 1\limits_\sim} $\end{document}) undergoes a unimolecular rearrangement to cis,cis-1,5-cyclooctadiene (\documentclass{article}\pagestyle{empty}\begin{document}$ {\mathop 2\limits_\sim} $\end{document}) and bimolecular formation of dimers \documentclass{article}\pagestyle{empty}\begin{document}$ {\mathop 3\limits_\sim}-{\mathop 5\limits_\sim} $\end{document} $\end{document}. The Arrhenius parameters are EA = 135.7 ± 4.4 kJ mole?1 and log(A/sec?1) = 12.9 ± 0.6 for the first reaction and EA = 66.1 ± 6.0 kJ mole?1 and log[A/(liter mole?1 sec?1)] = 5.5 ± 0.8 for the second reaction. Using thermochemical kinetics, the first reaction is shown to proceed via a rate determining Cope rearrangement of \documentclass{article}\pagestyle{empty}\begin{document}$ {\mathop 1\limits_\sim} $\end{document} to cis? 1,2-divinylcyclobutane (\documentclass{article}\pagestyle{empty}\begin{document}$ {\mathop 6\limits_\sim} $\end{document}), EA = 136.2 - 4.4 kJ mole?1 and log(A/sec?1) = 13.0 ± 0.6. The corresponding back reaction, \documentclass{article}\pagestyle{empty}\begin{document}$ {\mathop 6\limits_\sim}{\rightarrow}{\mathop 1\limits_\sim} $\end{document}, which was investigated separately, shows EA = 110.2 ± 1.2 kJ mole?1 and log(A/sec?1) = 10.9 ± 0.2. The heat of formation of \documentclass{article}\pagestyle{empty}\begin{document}$ {\mathop 6\limits_\sim} $\end{document} is determined to 188 ± 5.5 kJ mole?1. The mechanism of formation of dimers \documentclass{article}\pagestyle{empty}\begin{document}$ {\mathop 3\limits_\sim}-{\mathop 5\limits_\sim} $\end{document} is discussed. To allow the formal analysis of the kinetic problem, a simple algorithm to obtain the rate constants of competing first- and second-order reactions was developed.  相似文献   

16.
The kinetic feature of the anionic polymerization of N-PMI was investigated in THF. The polymerization system initiated with lithium tert-butoxide was revealed to be so-called “slow-initiation” system. The rate constant of the initiation reaction, ki, was obtained to be 4.2 × 10?3 (L mol?1 s?1) at ?72°C. The apparent rate constants of the propagation reaction, k, at ?72°C were individually obtained from each slope of the first-order plots in the later stages of the polymerizations for four different initiator concentrations. Each k is fairly close to that of initiation rate around 10?3. The propagation reaction was concluded to be dominated by ion-pair mechanism from the analysis of the kinetic data and the results of the addition effects of crown ether and common salt.  相似文献   

17.
2,2,4-Trimethyl-3-on-1-pentyl methacrylate (TMPM) was first synthesized from the condensation reaction of 2,2,4-trimethyl-1-pentanol-3-on with methacrylic acid. Second, the polymerization of TMPM and the copolymerization of TMPM with styrene (St) were carried out in benzene at 60°C, using 2,2′-azobisisobutyronitrile (AIBN) as an initiator. As the result of kinetic investigation, the rate of polymerization (Rp) could be expressed by: Rp = k[AIBN]0.5 [TMPM]1.0. Kinetic constants of polymerization of TMPM were obtained as follows: kp/k = 0.27 dm3/2 mole?1/2 sec?1/2, 2fkd = 1.23 × 10?5 sec?1, f = 0.73, Cm = 2.6 × 10?5, Cs = 1.1 × 10?5. From the results the reactivity of TMPM was found to be larger than that of methyl methacrylate. The overall activation energy was calculated to be 110 kJ mole?1. The following monomer reactivity ratios and Q, e values were obtained: TMPM(M1) ? St(M2): r1 = 1.50, r2 = 0.14, Q1 = 2.63, E1 = 0.45.  相似文献   

18.
The kinetics of the reversible reaction have been studied spectrophotometrically in acid solution under conditions in which both the forward and reverse reactions go to virtual completionand in which the reaction comes to a practical equlibrium. The rates of theforward (Rf) and reverse (Rr) reactions are given by where f, g, h, u, and v have the values (4 ± 1) × 10?5 mole/1.·s, (4.2 ± 0.2) × 10?5 mole2/1.2·s, (5.0 · 0.3) × 10?7 mole3/1.3·s, (1.1 ± 0.1) × 10?3 1.2/mole2·s, and (3.7 ± 0.2) × 10?3 1.3/mole3·s at 298.2°K and at an ionic strength of 2.00M maintained by adding sodium chloride. The stoichiometric equilibrium constant under similar conditions is 0.022 ± 0.003. Differentvalues of these parameters were obtained when sodium perchlorate and sodiumnitrate were used to control ionic strength. The results are compared with those from previous reports and a mechanism is proposed based upon an initial rapid equilibrium followed by a rate-determining attack of water upon H3AsO3I+, H2AsO3I, and HAsO3I?.  相似文献   

19.
Although the reaction of benzene with aluminum chloride has been quite thoroughly examined by prior investigators, the present report is the first one on formation of poly-p-phenylene in this system. Optimum conditions, which gave low yields, involved 7 days at 49–51deg;C. The presence of oxygen (oxidant) and presumably water (cocatalyst) was necessary in order for polymerization to occur. Physical and chemical properties, e.g., behavior toward Br2 and H2O2–CH3CO2H, indicate that the polymer structure is slightly different from that of the material from C6H6–AlCl3–CuCl2. The polymer from C6H6–AlCl3 may possess a lower molecular weight and exhibits a greater degree of structural irregularity in the form of dihydrobenzene, p-quinoid, or polynuclear regions. In the chemical studies, various compounds were used as models (tetrahydroquaterphenyl, triphenylene, and lower p-polyphenyls).  相似文献   

20.
Adducts and Salts Formed by Sulphurchlorides with AlCl3 The instability of the adduct 2 S2Cl2 · AlCl3 is proven. S2Cl2 · AlCl3 and S2Cl2 · 2 AlCl3 reported in the literature could not be found under proper conditions, their formation seems improbable. The product 2 SCl4 · 3 AlCl3, obtained by the reaction of [SCl3]+[AlCl4]? with elementary sulphur, is characterized as a double salt [SCl3]2+[AlCl4]? [Al2Cl7]?. The [Al2Cl7]? anion is also found as an intermediate during the thermal decomposition of [SCl3]+[AlCl4]? and when metallic aluminium reacts directly with S2Cl2. For SCl2 · AlCl3, the ionic character with a chlorsulfenium cation [SCl]+ is proven spectroscopically.  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