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 15 毫秒
1.
The 13C n.m.r. spectra of forty alkoxysilanes of the general type XnSi(OR)4–n (X = CH3, C6H5, H; R = CH3, C2H5, n-C3H7, i-C3H7, n-C4H9, i-C4H9, s-C4H9, n-C5H11, CH(CH3)(C6H5), C6H5) have been recorded and assigned. The chemical shifts of the α-carbon resonances of the alkoxy groups are shown to depend on both the nature of the alkoxy group and the number and type of substituents on the silicon. Regression analyses of the data give empirical substituent chemical shift (SCS) parameters for the silyl substituents. The β-carbon resonances are shown to be dependent on the presence of the silyl group, but not the specific silyl substituents.  相似文献   

2.
Novel copolymers of trisubstituted ethylene monomers, ring-substituted 2-phenyl-1,1-dicyanoethylenes, RC6H2CH=C(CN)2 (where R is 4-C6H5O, 2-C6H5CH2O, 3,4-(C6H5CH2O)2, 2-C6H5CH2O-3-CH3O, 3-C6H5CH2O-4-CH3O, 2-Cl-6-NO2, 4-Cl-3-NO2, 5-Cl-2-NO2) and 4-fluorostyrene were prepared at equimolar monomer feed composition by solution copolymerization in the presence of a radical initiator (ABCN) at 70°C. The composition of the copolymers was calculated from nitrogen analysis, and the structures were analyzed by IR, 1H and 13C-NMR. The order of relative reactivity (1/r 1) for the monomers is 3,4-(C6H5CH2O)2(31.0) > 2-C6H5CH2O-3-CH3O (24.8) > 3-C6H5CH2O-4-CH3O (15.2) > 4-C6H5O (3.1) > 4-Cl-3-NO2 (1.9) > 2-Cl-6-NO2 (1.6) > 5-Cl-2-NO2 (1.5) > 2-C6H5CH2O (1.4). High Tg of the copolymers, in comparison with that of poly(4-fluorostyrene) indicates a substantial decrease in chain mobility of the copolymer due to the high dipolar character of the trisubstituted ethylene monomer unit. Decomposition of the copolymers in nitrogen occurred in two steps, first in the 270-400°C range with residue, which then decomposition in 400–800°C range.  相似文献   

3.
Novel trisubstituted ethylenes, ring-substituted butyl 2-cyano-3-phenyl-2-propenoates, RPhCH=C(CN)CO2C4H9 (where R is 2-C6H5CH2O, 3-C6H5CH2O, 4-C6H5CH2O, 4-CH3COO, 3-CH3CO, 4-CH3CO, 4-CH3CONH, 2-CN, 3-CN, 4-CN, 4-(CH3)2N, 4-(C2H5)2N) were prepared and copolymerized with styrene. The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring-substituted benzaldehydes and butyl cyanoacetate, and characterized by CHN analysis, IR, 1H and 13C-NMR. All the ethylenes were copolymerized with styrene (M1) in solution with radical initiation (ABCN) at 70°C. The compositions of the copolymers were calculated from nitrogen analysis and the structures were analyzed by IR, 1H and 13C-NMR. The order of relative reactivity (1/r1) for the monomers is 4-C6H5CH2O (6.39) > 2-C6H5CH2O (2.06) > 3-CH3CO (1.86) > 3-C6H5CH2O (1.78) > 4-CH3COO (1.58) > 3-CN (1.47) > 4-CN (1.21) > 4-(C2H5)2N (1.19) > 4-(CH3)2N (1.18) > 2-CN (1.04) > 4-CH3CO (0.71) > 4-CH3CONH (0.63). Decomposition of the copolymers in nitrogen occurred in two steps, first in the 200–500°C range with residue (3.6–9.5% wt), which then decomposed in the 500–800°C range.  相似文献   

4.
The behaviour under electron impact (70 eV) which includes some rearrangement processes of some tetraorganodiphosphanedisulfides R2P(S)-P(S)R2 (R ? CH3, C2H5, n-C3H7, n-C4H9, C3H5, C6H5) and CH3RP(S)–P(S)CH3R (R ? C2H5, n-C3H7, n-C4H9, C6H5, C6H5, C6H5,CH2) is reported and discussed. Fragmentation patterns which are consistent with direct analysis of daughter ions and defocusing metastable spectra are given. The atomic composition of many of the fragment ions was determined by precise mass measurements. In contrast to compounds R3P(S) loss of sulphur is not a common process here. The first step in the fragmentation of these compounds is cleavage of one P–C bond and loss of a substituent R?. The second step is elimination of RPS leading to [R2PS]+ from which the base peaks in nearly all the spectra arise. The phenyl substituted compounds give spectra with very abundant [(C6H5)3P]+. and [(C6H5)2CH3P]+. ions respectively, resulting from [M]+. by migration of C6H5. Rearrangement of [M]+. to a 4-membered P-S ring system prior to fragmentation is suggested.  相似文献   

5.
Effects of steric crowding of the substituent of carboxylate counteranions on living cationic polymerization of isobutyl vinyl ether (IBVE) were investigated with the use of two series of carboxylic acids with various carbonyl substituents [RCOOH; R = (aliphatic series) CH3CH2, (CH3)2CH, (CH3)3C; (aromatic series) C6H5CH2, (C6H5)2CH, (C6H5)3C] in conjunction with tin tetrabromide (SnBr4) and 1,4-dioxane (DO) in toluene at 0°C. The overall polymerization rate increased with increasing the bulkiness of the substituents R in both the series: R = CH3 (1) ≃ CH3CH2 (1) < (CH3)2CH (1.76) < (CH3)3C (2.31); C6H5CH2 (0.84) < (C6H5)2CH (0.98) < (C6H5)3C (1.74); the values in the parentheses show the relative polymerization rate. In all the polymerizations, the number-average molecular weight (Mn) of the polymers was directly proportional to monomer conversion and in good agreement with the calculated values, assuming that one RCOOH molecule forms one polymer chain. The living nature of these polymerizations was further confirmed by a linear increase in Mn of the polymers upon sequential addition of a fresh monomer feed to the almost completely polymerized reaction mixtures. In the polymerizations with sterically less hindered carboxylic acids [R = CH3CH2, (CH3)2CH, C6H5CH2, (C6H5)2CH], the molecular weight distribution (MWD) of the polymers was very narrow (Mw/Mn < 1.1) throughout the polymerizations. In contrast, with bulkier substituent-containing counterparts [R = (CH3)3C, (C6H5)3C], the polymerizations led to the polymers of relatively broad MWD (Mw/Mn ≅ 1.5 at ca. 100% monomer conversion). The bulky substituents such as (CH3)3C and (C6H5)3C may decrease the interconversion rate between a dormant and an active species and increase the time-average concentration of the active growing species. The stereoregularity of the obtained polymers was not changed much with the steric environment of the counteranion (meso: 66–69%). © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2923–2932, 1999  相似文献   

6.
Novel copolymers of trisubstituted ethylene monomers, ring-substituted 2-phenyl-1,1-dicyanoethylenes, RC6H4CH = C(CN)2 (where R is 2-F, 2-CN, 3-CN, 4-CN, 3-C6H5O, 4-C6H5O, 2-C6H5CH2O, 3-C6H5CH2O, 4-C6H5CH2O, 4-CH3CO2, 4-CH3CONH, 4-(CH3)2N) and styrene were prepared by solution copolymerization in the presence of a radical initiator (ABCN) at 70°C. The composition of the copolymers was calculated from nitrogen analysis, and the structures were analyzed by IR, 1H and 13C-NMR. The order of relative reactivity (1/r 1) for the monomers is 4-(CH3)2N (3.35) > 4-C6H5CH2O (3.1) > 2-C6H5CH2O (1.77) > 3-C6H5CH2O (1.72) > 4-C6H5O (1.70) > 4-CH3CO2 (1.58) > 2-F (1.11) > 3-C6H5O (0.90) > 3-CN (0.88) > 2-CN (0.86) > 4-CH3CONH (0.84) > 4-CN (0.76). Relatively high Tg of the copolymers in comparison with that of polystyrene indicates a decrease in chain mobility of the copolymer due to the high dipolar character of the trisubstituted ethylene monomer unit. Decomposition of the copolymers in nitrogen occurred in two steps, first in the 200–500°C range with residue (1–10% wt), which then decomposed in the 500–800°C range.  相似文献   

7.
The pyrolysis of 2% CH4 and 5% CH4 diluted with Ar was studied using both a single–pulse and time–resolved spectroscopic methods over the temperature range 1400–2200 K and pressure range 2.3–3.7 atm. The rate constant expressions for dissociative recombination reactions of methyl radicals, CH3 + CH3 → C2H5 + H and CH3 + CH3 → C2H4 + H2, and for C3H4 formation reaction were investigated. The simulation results required considerably lower value than that reported for CH3 + CH3 → C2H4 + H2. Propyne formation was interpreted well by reaction C2H2 + CH3P-C3H4 + H with ?? = 6.2 × 1012 exp(?17 kcal/RT) cm3 mol?1 s?1.  相似文献   

8.
The Flory–Huggins interaction parameters χ for 23 gases (He, Ne, Ar, Kr, Xe, H2, N2, O2, N2O, CO2, CH4, C2H4, C2H6, C3H6, C3H8, 1,3-C4H6, four C4H8's, n-C4H10, iso-C4H10, and n-C5H12) in five rubbery polymers (1,2-polybutadiene (PB), poly(ethylene-co-vinyl acetate)) (EVAc), polyethylene (PE), polypropylene (PP), and poly(dimethyl siloxane) (PDMS) were determined from either literature data on Henry's law coefficient and partial molar volume or those on sorptive dilation for each polymer/gas system. Values of χ for the gases increased in the order of PDMS < PP ≡ PB < EVAc ≡ PE. Among the gases except He and H2 whose χ values are not reliable, Ne and Xe have respectively the highest and the lowest values of χ for the polyolefins. The χ values of the hydrocarbons were compared together with previously reported χ values of n-alkanes C3-C10. The dependencies of χ upon concentration and temperature were discussed on the basis of the literature data. © 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35 : 1049–1053, 1997  相似文献   

9.
Liquid–liquid (LL) critical demixing loci have been experimentally determined in the (T,P) projection for some polystyrene/solvent systems with nonspecific interaction for (∼ 270 K < T < ∼ 500 K) and (0 MPa < P < 200 MPa). A lower homogeneous double critical pressure and lower homogeneous double critical temperature have been located for a solution of PS (Mw = 2.0 × 106) dissolved in an n-heptane/methylcyclohexane mixture [PS/n-C7H16/CH3C6H11//0.029/0.194/0.777 (wt. fractions)]. That solution forms the first example of a polymer/solvent with nonspecific interaction that exhibits two double critical points. A symmetrical representation of the LL critical loci in the (P,T) plane is developed. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 2747–2753, 1999  相似文献   

10.
The kninetics of acid-catalyzed acetalization and ketalization of poly(vinyl alcohol) (PVA) were systematically studied in completely homogeneous media with carefully selected solvents. Thus the acetalization reaction was run in water with six aldehydes [R1CHO (R1 = H, CH3, C2H5, n-C3H7, i-C3H7, ClCH2)], whereas the ketalization in dimethylslfoxide with 11 ketones [R2CH3CO (R2 = CH3, C2H5, n-C3H7, i-C3H7, n-C4H9, i-C4H9, tert-C4H9, C6H5CH2, C6H5CH2CH2), cyclopentanone, and cyclohexanone]. The latter was difficult to proceed in aqueous media. Both reactions were reversible and bimolecular and, despite the use of different solvents, gave similar heats of reaction (7.5 kcal/mol) and activation energies (ca. 15 kcal/mol) except for the case of formaldehyde and chloroacetaldehyde; however the equilibrium constants at 25°C showed that the acetalization is thermodynamically much more favored than the ketalization (ca. 5000 vs. 0.01–0.9), probably because of steric hindrance of the ketone substrate. The rate constants of hydrolysis (reverse reactions) for the poly(vinyl acetal) and poly(vinyl ketal) followed the Hammett-Taft equation to give a single p* (=3.60) that is very close to that for the hydrolysis of diethyl acetal and ketal. From these and other data, it was concluded that the polymer hydrolysis, as well as PVA acetalization and ketalization, are all electrophilic reaction where the formation of hemiacetal or hemiketal is the rate-determining step. © 1996 John Wiley & Sons, Inc.  相似文献   

11.
The effect of organic sulfur compounds on the radical polymerization of methyl methacrylate initiated by azobisisobutyronitrile at 50°C. has been studied. The sulfur compounds used were benzene-type polysulfides (C6H5CH2? Sn? CH2C6H5; n = 0–4), benzyl mercaptan, and sulfur (S8). All sulfur compounds studied, except dibenzyl, dibenzyl monosulfide, and dibenzyl disulfide, were found to behave as retarders under these experimental conditions. Chain-transfer constants of these compounds were determined from rate measurements and from the conventional method based on numberaverage degree of polymerization. Chain-transfer constants of benzyl-type polysulfides were less than those of mercaptan and sulfur and increased with increasing sulfur. The correlation of the reactivities of sulfur compounds as transfer agents and their molecular structures is discussed.  相似文献   

12.
In the reaction of [C5H5Mn(CO)2(NO)] [X] ([X] = [BF4], [PF6]) with p-substituted triarylphosphines P(p-C6H4?Y)3 [Y = CF3, Cl, F, C6H5, CH3, OCH3, N(CH3)2] the asymmetric monosubstitution products [C5H5Mn(CO)(NO)P(p-C6H4?Y)3] [X] are formed, which can be converted into the neutral esters C5H5Mn(COOC10H19)(NO)P(p-C6H4?Y)3 by natrium menthoxide. The diastereoisomers (+)579? and (?)579?C5H5Mn(COOC10H19)(NO)P(p-C6H4?Y)3 are separated by fractional crystallisation and transformed into the enantiomeric salts (+)579? and (?)579-[C5H5Mn(CO)(NO)P(p-C6H4?Y)3] [X] by cleavage with HCl and precipitation with NH4PF6. The (+)579? and (?)579? rotating salts in the reaction with LiC6H5 yield the carbonyl addition products (+)579? and (?)579? C5H5Mn(COC6H5)(NO)P(p-C6H4?Y)3 and the ring addition products (+)579? and (?)579?(exo-C6H5)C5H5Mn(CO)(NO)P(p-C6H4?Y)3, which can be separated by chromatography.The salts (+)579? and (?)579?[C5H5Mn(CO)(NO)P(p-C6H4?Y)3] [X] and the cyclopentadiene complexes (+)579? and (?)579-(exo-C6H5)C5H5Mn(CO)(NO)P(p-C6H4?Y)3 are configurationally stable, whereas the esters (+)579? and (?)579?C5H5Mn(COOC10H19)(NO)P(p-C6H4?Y)3 and the benzoyl complexes (+)579? and (?)579?C5H5Mn(COC6H5)(NO)P(p-C6H4?Y)3 epimerise or racemise in solution.The rate of racemisation of the benzoyl compounds (+)579? and (?)579C5H5Mn(COC6H5)(NO)P(p-C6H4?Y)3 was measured polarimetrically in the temperature range 0–45° C. It turned out that electron-releasingsubstituents Y in the ligand P(p-C6H4?Y)3 increase the half-lives, whereas electron-attracting substituents decrease the half-lives. There is a linear correlation between the σ-constants of the substituents and the rate constants of the racemisation (reaction constant p = +2.14).  相似文献   

13.
Earlier work has shown that the cyclo-addition of a ketene to a conjugated diene is always (a) 2 + 2, (b) polarily directed, and (c) suprafacial with respect to the diene C?C. The adducts of ketenes and cyclopentadiene are thus always 7-substituted bicyclo [3.2.0] hept-2-ene-6-ones. New evidence is presented to show that unsymmetrically substituted ketenes add to cyclopentadiene in such a manner that the larger substituent has a greater tendency to take up the endo- position in the adduct. This is interpreted to mean that a ketene participates in such reactions antarafacially. Thus the ketene approaches cyclopentadiene (a) with its functional plane perpendicular to that of the ring, (b) with the carbonyl carbon over the middle of the ring, and (c) with the larger of the two substituents oriented preferentially away from the ring (transition state 11 ). This endo-specificity for the larger ketene substituent is demonstrated by the indicated endo/exo ratios observed in the cyclo-adducts from ketenes with the following substituent pairs: C6H5/H = >95/<5, CH3/H = 98/2, Cl/H = 97/3, CH3O/H = >95/<5, C6H5/CI = >95/<5, C6H5/CH3 = >95/<5, CH3/Cl = 80/20, CH3/CH?CH2 = ~65/35, C2H5/CH3 = ~60/40, n-C3H7/CH3 = ~60/40, CH3/Br = 56/44. These ratios enable a list to be compiled indicating the endo-specificity of the ketene substituents. The order closely parallels the space filling capacity as derived by other methods. The establishment of such ratios required reliable configurational assignments at carbon 7. These were derived by five methods based upon the following effects: (1) Both H? C7 and CH3? C7 cause nmr. signals at higher field in endo-position (compared with exo). (2) The CH3? C7 group in exo-position gives rise to a nuclear Overhauser effect with the vicinal H? C1, and in one case also with the trans-annular H? C5. (3) The nmr.-coupling constants of H? C7-exo (observed at H? C7) with H? C1 is always larger than of H? C7-endo. (4) The coupling constant of H? C7-exo with H? C6 (known to be exo-) of the LiAlH4 reduction products of the cyclo-adducts (observed at H? C6) is always larger than that of H? C7-endo. (5) The nmr. signals of most protons in the cyclo-adducts are at higher field in benzene than in chloroform solution; this “benzene shift” is larger for H? C7 or for CH3? C7 when in exo- than when in endo-position.  相似文献   

14.
The photochemical reaction of π-C5H5Fe(CO)(CNC6H11)COCH3 (I) gave the heterocyclic compound π-C5H5Fe(CO)[(C=NC6H11)2(CH3)] (II) involving N-coordination to the iron atom. The analogous complex is obtained by the photo-induced reaction of π-C5H5Fe(CO)2CH3 with C6H11NC. A similar reaction of π-C5H5Fe(CO)[CNC(CH3)3]CH3 with C6H11NC gave π-C5H5Fe(CO)[(C=NC6H11) {C=N(CH3)3}(CH3)] (IV) involving different N-substituted imino groups. The possible pathways leading to formation of II are discussed. The mass spectra of these complexes were also investigated.  相似文献   

15.
Five-membered cyclic esters of phosphoric acid of the general formula: ? CH2CH(R)OP(O)-(OR′)O? polymerize readily to solid, soluble polymers of high molecular weight without any rearrangement known for various tri- and pentavalent organophosphorus monomers. 1H-, 13C-, and 31P-NMR spectra of polymers confirmed their linear structure: where R is H, with R′ = CH3, C2H5, n-C3H7, i-C3H7; n-C4H9, CCl3CH2, or C6H5, or R is CH2Cl and R′ is C2H5. The use of n-C4H9Li, (C5H5)2Mg, or (i-C4H9)3Al as initiators leads to polymers with M n = 104–105.  相似文献   

16.
Poly(2,3-dialkylbutanediol-1,4 terephthalates) with the alkyl substituents CH3, C2H5, n-C3H7, iso-C3H7, n-C4H9, and n-C10H21, andn-C16H33 were synthesized from the corresponding 2,3-dialkylbutanediols-1,4 and dimethyl terephthalate or terephthaloyl chloride. The substituents of the butanediol-1,4 portion of the polyterephthalates influence the 13C NMR chemical shifts of the carbon atoms near the branching site, the glass transition (Tg), and the crystallizability. Small alkyl substituents do not change the Tg of the polymers, whereas bulky substituents such as the isopropyl group increase the Tg and long normal alkyl groups as substituents decrease the Tg of the polymers. Crystallinity in these polyterephthalates was found only with CH3 and C16H33 as the 2,3-dialkyl substituents in the butanediol-1,4 portion of the polyester. This crystallinity of polyterephthalate of 2,3-di-C16H33 substituted butanediol-1,4 could be assigned to side-chain crystallization of the paraffinic groups.  相似文献   

17.
1,3-Butadiene (1,3-C4H6) was heated behind reflected shock waves over the temperature range of 1200–1700 K and the total density range of 1.3 × 10−5 −2.9 × 10−5 mol/cm3. Reaction products were analyzed by gas-chromatography. The concentration change of 1,3-butadiene was followed by UV kinetic absorption spectroscopy at 230 nm and by quadrupole mass spectrometry. The major products were C2H2, C2H4, C4H4, and CH4. The yield of CH4 for a 0.5% 1,3-C4H6 in Ar mixture was more than 10% of the initial 1.3-C4H6 concentration above 1500 K. In order to interpret the formation of CH4 successfully, it was necessary to include the isomerization of 1,3-C4H6 to 1,2-butadiene (1,2-C4H6) and to include subsequent decomposition of the 1,2-C4H6 to C3H3 and CH3. The present data and other shock tube data reported over a wide pressure range were qualitatively modeled with a 89 reaction mechanism, which included the isomerizations of 1,3-C4H6 to 1,2-C4H6 and 2-butyne (2-C4H6). © 1996 John Wiley & Sons, Inc.  相似文献   

18.
The mass spectra of several alkyl phenyl tellurides, C6H5TeR (R = CH3, CD3, C2H5, n-C3H7, i-C3H7 and n-C4H9) have been studied with special emphasis on the fragmentation patterns involving cleavage of the alkyl and aryl tellurium–carbon bonds. Each compound exhibited intense parent ions. The rearrangement ions [C6H6Te]+? and [C6H6]+? were found in the spectra of phenyl ethyl and higher tellurides. Two other rearrangement ions [HTe]+ and [C7H7]+ were observed in the spectrum of each compound. Examination of the mass spectrum of phenyl methyl-d3 telluride demonstrated that the [HTe]+ ions derive hydrogen from the phenyl group.  相似文献   

19.
Electrophilic trisubstituted ethylene monomers, ring-substituted methyl 2-cyano-3-phenyl-2-propenoates, RC6H4CH?C(CN) CO2CH3 (where R is 4-C2H5O, 4-C3H7O, 4-C4H9O, 3-C6H5O, and 3-CN), were prepared and copolymerized with styrene. The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring-substituted benzaldehydes and methyl cyanoacetate, and characterized by CHN elemental analysis, IR, 1H- and 13C-NMR. All the propenoates were copolymerized with styrene (M1) in solution with radical initiation (AIBN) at 70°C. The compositions of the copolymers were calculated from nitrogen analysis and the structures were analyzed by IR, 1H- and 13C-NMR. The order of relative reactivity (1/r1 ) for the monomers is 3-CN (1.21) > 3-C6H5O (1.16) > 4-C2H5O (0.94) > 4-C3H7O (0.8305) > 4-C4H9O (0.616). The high T g's of the copolymers (> 130°C) in comparison with that of polystyrene indicate a substantial decrease in the chain mobility of the copolymers due to the high dipolar character of the trisubstituted monomer unit. Gravimetric analysis indicated that the copolymers decompose in the range 300–400°C.  相似文献   

20.
The cation [CpRu(η6-C10H8)]+ was shown to exchange naphthalene for other arenes under visible-light irradiation to form the complexes [CpRu (η6-arene)]+ (arene = C6H6, 1,4-C6H4Me2, 1,3,5-C6H3Me3, or 1,2,4,5-C 6H2Me4) in 70–95% yields. The reaction rate of exchange decreases in the series arene = 1,4-C6H4Me2 > C6H6 > 1,3,5-C6H3Me3 > 1,2,4,5-C 6H2Me4 >> C6Me6 and increases with the coordinating ability of the solvent in the order CH2Cl2 < THF—CH2Cl2 mixture (1: 1) < acetone.  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