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 aminolysis reactions of O-ethyl S-(Z-phenyl) dithiocarbonates (Z=p-CH3, H, p-Cl, and p-NO2) with anilines (AN) and N,N-dimethylanilines (DMA) in acetonitrile at 30.0°C are investigated. Relatively small values of βXnuc,0.4 ca. 0.7) and βZlg −0.1 ca. −0.4) for both ANs and DMAs, significantly large kH/kD values (1.1 ca. 1.9) involving deuterated anilines, and large negative ρXZ values for ANs (−0.56) are interpreted to indicate a concerted mechanism for both ANs and DMAs but with a hydrogen bonded four-center type transition state (TS) for ANs. The relative leaving ability, k(Z=p-NO2)/k(Z=p-CH3), is smaller for ANs than for DMAs, especially for a weaker nucleophile (1.9 and 4.7 for AN and DMA, respectively, with X=p-Cl). This suggests that the rate enhancement by the hydrogen-bond formation in the four-center type TS for AN is greater for a weaker nucleofuge (Z=p-CH3), especially when the nucleophile (X=p-Cl) is weaker. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 419–423,1998  相似文献   

2.
On the Base-Catalysed H/D-Exchange of the Acetylenic Hydrogen Atom in Aromatic Alkynyl Compounds H/D-exchange rates for a number of compounds of the general type 1 (X = p-CH3O, m-CH3O, p-CF3, m-CF3, p-CH3, p-Cl, H; Z ? O, NH, CH2) were determined in N-methyl-pyrrolidine (NMP)/D2O mixtures at 25° (see Table 1). It is shown that the log k values of the H/D-exchange correlate nicely (r = 0.995) with the chemical shift of the acetylenic proton in 1 . Thus, the H/D-exchange rate is given by log k (min?1) = 2.91 · δ (ppm) - 7.79 for the NMP/D2O mixture at 25°.  相似文献   

3.
Hydrolysis of the imine function of a series of Schiff bases derived from pyrrole-2-carboxaldehyde and substituted anilines (X = H, p-OCH3, p-OC2H5, p-CH3, p-Cl, p-Br, m-CN, m-NO2, p-NO2) was studied in all of the pH ranges. The hydrolysis curves log kabs (mn?1) = f(pH) were established in buffered aqueous methanol by polarography or amperometry. The shapes of the curves obtained for pH > 5 indicates that N-pyrrolylmethylene-2 anilines hydrolyse according to the same mechanism as N-benzylidene anilines. The particular stability of these products for pH < 5 permits one to obtain complete hydrolysis curves in acid media. A very good Hammett correlation (kobs = + 1,73) has been established from the maxima which appear for strong acidity. This stability is interpreted as due to the specific electron-donating effect of the pyrrole nucleus. The influence of the structural parameter X on the morphology of the curves log kabs = f(pH) and on the hydrolysis mechanism of the imine function is discussed in all of the pH ranges.  相似文献   

4.
The mass spectrometry of substituted benzenesulfonylhydrazides (X.C6H4.SO2NHNH2) has been studied, with X = p-CH3, H, p-CH3O and p-Br. The intensities of [X? C6H4SO2]+ and [X? C6H4]+ follow the Hammett relationship [In(Z/Z0) =ρδp+] with ρ of 0.375 and 2.37, respectively.  相似文献   

5.
The kinetics and mechanism of the nucleophilic substitution reactions of p‐chlorophenyl aryl chlorophosphates ( 2 ) with anilines are investigated in acetonitrile at 55°C. Relatively large magnitudes of ρX and βX values are indicative of a large degree of bond making in the TS. Smaller magnitudes of ρX (0.20 for X = H) and ρXY (?0.30) than those for the corresponding reactions with phenyl aryl chlorophosphates ( 1 ) (ρX = 0.54 for X = H and ρXY = ?1.31) are interpreted to indicate partial electron loss, or shunt, towards the electron acceptor equatorial ligand (p‐ClC6H4O‐) in the bipyramidal pentacoordinated transition state. The inverse secondary kinetic isotope effects (kH/kD = 0.64–0.87) involving deuterated aniline (ND2C6H4X) nucleophiles, and small ΔH? and large negative ΔS? are obtained. These results are consistent with a concerted nucleophilic substitution mechanism. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 632–637, 2002  相似文献   

6.
Synthesis of Fluoro-λ5-monophosphazenes and Fluoro-1,3-diaza-2λ5,4λ5-diphosphetidines by Means of the Staudinger Reaction 35 Tetrafluoro- and 2 difluorodiaza-diphosphetidines as well as 4 difluoro- and 30 monofluoro-λ5-monophosphazenes were prepared by the Staudinger reaction between tervalent phosphorus fluorides, RnPF3?n (n = 1, 2; R = R2N, (CH2)5N, O(CH2)4N, RO, (CH2O)2, alkyl, aryl) and phenylazides, X? C6H4N3 (X = H, 4-CH3, 4-Cl, 4-Br, 4-NO2, 3-NO2). PF3 does not react with phenylazide The influence of substituents on the structure of the reaction products is discussed. Kinetic measurements allowed to determine the constants λPI of the substituents (CH2)5N, O(CH2)4N and R(C6H5)N (R = CH3, C2H5, n-C4H9).  相似文献   

7.
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   

8.
The synthesis and characterisation of the complexes [(p-CH3C6H4NCH(C6H3Y))Pd(OAc)]2 (II) are reported. These complexes react at very different rates with carbon monoxide in methanol to give the ortho-substituted esters, p-CH3C6H4NCHC6H3Y - 2R, R = CO2CH3, with electron withdrawing Y substituents slowing the reaction. The 13C{1H} data for II show a linear correlation of δ(C(2)) in the 5′-complexes (Y trans to PdC) with δ(C(4)) of monosubstituted benzene compounds. For Y = 5′-NO2, 4′-NO2 and 4′-Cl, the bis complex [{p-CH3C6H4NCH(
is formed in a secondary reaction.  相似文献   

9.
We report the first example of aryl hydrogen scrambling occurring in a molecular anion prior to or accompanying fragmentation, i.e. for the reaction [M]?· → [M ? H2NO2.]? from o-NO2? C6H4? X? C6H5 (X = O or S). Proximity effects occur in these spectra when X = CO, NH, O, or S, and certain of these have been substantiated by 2H and 18O labelling.  相似文献   

10.
Electrophilic trisubstituted ethylenes, ring-disubstituted ethyl 2-cyano-3-phenyl-2-propenoates, RPhCH?C(CN)CO2C2H5 (where R is 3-Br-4-CH3O, 5-Br-2-CH3O, 3-F-2- CH3, 3-F-4-CH3, 4-F-2-CH3, 4-F-3-CH3, 5-F-2-CH3, 2-Cl-5-NO2, 2-Cl-6-NO2, 4-Cl-3- NO2) were prepared and copolymerized with styrene. The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring-disubstituted benzaldehydes and ethyl 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 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 5-Br-2-CH3O (1.02) > 4-Cl-3-NO2 (0.93) > 3-F-4-CH3 (0.81) > 2-Cl-6-NO2 (0.77) > 2-Cl-5-NO2 (0.71) > 3-Br-4-CH3O (0.66) > 4-F-3-CH3 (0.60) > 3-F-2-CH3 (0.38) > 4-F-2-CH3 (0.31) > 5-F-2-CH3 (0.16). 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 250–500°C range with residue (2–26% wt.), which then decomposed in the 500–800°C range.  相似文献   

11.
Various p-substituted benzyl p-hydroxyphenyl methyl sulfonium salts ( 2 ) were synthesized and their initiator activities were evaluated in bulk polymerization of glycidyl phenyl ether (PGE). The order of the activity was found to be 2b (X = CH3) > 2a (X = H) ≈ 2c (X = Cl) > 2d (X = NO2), indicating that the introduction of an electron-donating group enhanced the activity. In Hammett's plots, the logarithm of the ratio of the polymerization rates (log kx/kH) was correlated with σ+ρ better than with σp and a negative ρ+ value (-1.18) was obtained. Reaction of 2a with benzyl mercaptan mainly gave dibenzyl sulfide and p-hydroxyphenyl methyl sulfide. The obtained results seemed to demonstrate that the OH group of the aryl group yielded no proton as initiator for the polymerization, whereas the benzyl group caused the polymerization, which was initiated by the corresponding benzyl cation formed by C? S bond cleavage. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
Pulsed gradient spin‐echo (PGSE) diffusion characteristics for a) the new [brucinium][X] salts 6 a – f [ a : X=BF4?; b : X=PF6?; c : X=MeSO3?, d : X=CF3SO3?; e : X=BArF?; f : X=PtCl3(C2H4)?], b) 4‐tert‐butyl‐N‐benzyl analogue, 7 and c) the aryl carbocations (p‐R‐C6H4)2CH 9 a (R=CH3O) and 9 b (R=(CH3)2N), (p‐CH3O‐C6H4)xCPh3?x+ 10 a – c (x=1–3, respectively) and (p‐R‐C6H4)3C+ 11 (R=(CH3)2N) and 12 (R=H) all in several different solvents, are reported. The solvent dependence suggests strong ion pairing in CDCl3, intermediate ion pairing in CD2Cl2 and little ion pairing in [D6]acetone. 1H, 19F HOESY NMR spectra (HOESY: heteronuclear Overhauser effect spectroscopy) for 6 and 7 reveal a specific approach of the anion with respect to the brucinium cation plus subtle changes, which are related to the anion itself. Further, for carbocations 9 – 12 , (all as BF4? salts) based on the NOE results, one finds marked changes in the relative positions of the BF4? anion. In these aryl cationic species the anion can be located either a) very close to the carbonium ion carbon b) in an intermediate position or c) proximate to the N or O atom of the p‐substituent and remote from the formally positive C atom. This represents the first example of such a positional dependence of an anion on the structure of the carbocation. DFT calculations support the experimental HOESY results. The solid‐state structures for 6 c and the novel Zeise's salt derivative, [brucinium][PtCl3(C2H4)], 6 f , are reported. Analysis of 195Pt NMR and other NMR measurements suggest that the η2‐C2H4 bonding to the platinum centre in 6 f is very similar to that found in K[PtCl3(C2H4)]. Field dependent T1 measurements on [brucinium][PtCl3(C2H4)] and K[PtCl3(C2H4)], are reported and suggested to be useful in recognizing aggregation effects.  相似文献   

13.
Zusammenfassung Die magnetische Nichtäquivalenz der Methylenprotonen in 2-R-Benzo-1,3-dioxanen, mit R=C6H5,o-ClC6H4,m-ClC6H4,p-ClC6H4,o-BrC6H4,m-BrC6H4,p-BrC6H4,o-NO2C6H4,m-NO2C6H4,p-NO2C6H4,o-CH3OC6H4,m-CH3OC6H4,p-CH3OC6H4, -Naphthyl, -Naphthyl, -Thienyl und -Thienyl, wurde in (CD3)2CO, CDCl3, CCl4, CS2 und C6D6 untersucht.
Magnetic non equivalent methylenic protons of some 2-substituted 1.3-benzodioxanes
The magnetic non equivalence of the methylenic protons in 2-R-1.3-benzodioxanes, with R=C6H5,o-ClC6H4,m-ClC6H4,p-ClC6H4,o-BrC6H4,m-BrC6H4,p-BrC6H4,o-NO2C6H4,m-NO2C6H4,p-NO2C6H4,o-CH3OC6H4,m-CH3OC6H4,p-CH3OC6H4, -naphthyl, -naphthyl, -thienyl and -thienyl were studied in (CD3)2CO, CDCl3, CCl4, CS2 and C6D6.
  相似文献   

14.
Reaction of [Au(DAPTA)(Cl)] with RaaiR’ in CH2Cl2 medium following ligand addition leads to [Au(DAPTA)(RaaiR’)](Cl) [DAPTA=diacetyl-1,3,5-triaza-7-phosphaadamantane, RaaiR’=p-R-C6H4-N=N- C3H2-NN-1-R’, (1—3), abbreviated as N,N’-chelator, where N(imidazole) and N(azo) represent N and N’, respectively; R=H (a), Me (b), Cl (c) and R’=Me (1), CH2CH3 (2), CH2Ph (3)]. The 1H NMR spectral measurements in D2O suggest methylene, CH2, in RaaiEt gives a complex AB type multiplet while in RaaiCH2Ph it shows AB type quartets. 13C NMR spectrum in D2O suggest the molecular skeleton. The 1H-1H COSY spectrum in D2O as well as contour peaks in the 1H-13C HMQC spectrum in D2O assign the solution structure.  相似文献   

15.
V. Gani  P. Viout 《Tetrahedron》1978,34(9):1333-1336
Micellar effects of CTAB upon the alkaline hydrolysis of CF3-CO-N(CH3)C6H5, CHCl2-CO-N(CH3)C6H4X and CH2Cl-CO-N(CH3)C6H4X, (X=p-OCH3H,p-Cl) are reported. Variations of kobs, and of kinetic order of the reaction with respect to HO? ion, are interpreted as an acceleration of HO?-catalyzed steps, and a decrease of catalysis by water for decomposition of tetrahedral intermediates; these two effects oppose each other in HO? and H2O catalyzed steps. Differences between micellar and DMSO effects suggest a very small local concentration of HO? ions in micelles.  相似文献   

16.
A study of the thermal decomposition of an acetylene–ethane-d6 mixture indicates that the rate constant for hydrogen abstraction from acetylene by methyl is more than 20 times less than for abstraction from ethane. Isotopic exchange is initiated by a rapid reaction between product D atoms and C2H2. A series of experiments involving the reactions of a D2–acetylene mixture indicated that a molecular exchange process was also occurring, and it was shown that d[C2HD]/dt = k[D2]0.7[C2H2]0.3, effective activation energy = 15.8 kcal/mol. This mechanism made an insignificant contribution to isotope exchange in C2H2–C2D6 mixtures.  相似文献   

17.
Analysis of the dipole moments of N-trimethylammoniobenzamidates Me3N+-N?COC6H4X, with X = H, p-F, p-Cl or p-NO2, and of N-aroyliminodimethylsulphur(IV) Me2SNCOC6H4X (X = H and p-NO2) shows that, as solutes, these compounds exist in the syn conformation. Models are proposed for N-trimethylammonio-orthochloroben-zamidate and N-orthocyanobenzoyliminodimethylsulphur(IV). The (Me3N+-Nt-) and (SN) dipole moments, and the (N-CO) and (SN-CO) mesomeric moments, are derived and discussed.  相似文献   

18.
Acetylmethyltriphenylarsonium bromide 6 in the presence of potassium carbonate and trace water reacted with 2,2-dimethyl-l,3-dioxa-5-substituted-benzylidene-4,6-dione 2 at room temperature to give cyclopropane derivatives cis-l-acetyl-2-aryl-6,6-dimethyl-5,7-dioxospiro-[2,5]-4,8-octadiones 7 (X=p-CH3, p-Cl, H, p-NO2) or β,γ-trans-β-acetyl-γ-aryl-γ-butyrolactones 8 (X=p-CH3O, p-N(CH3)2, 3′,4′-OCH2O-) with good yield and high stereoselectivity.  相似文献   

19.
The velocity of the hydrogen ion catalysed hydrolysis of p-nitrophenyl-diazo-methane (I) has been measured in H2O? D2O mixtures, giving an isotopic αi = 0.49. The product isotope effect r = 5.1, determined from product analyses, combined with the (overall) solvent isotope effect kH/kD = 2.81, yields the primary kinetic isotope effect (kH/kD)I = 3.8, and the secondary kinetic isotope effect (kH/kD)II = 0.75. The CICH2COOH-catalysed hydrolysis of I in H2O? D2O mixtures gave a straight-line plot of kn/kH versus the atomic fraction n of deuterium. With four carboxylic acids, as catalysts, values of about 4.3 for the kinetic (overall) isotope effects were observed.  相似文献   

20.
The kinetics of dimerization of arylmercurials XC6H4HgCl (X ? H, p-CH3, m-CH3, p-Cl, m-OCH3, m-CO2C2H5, and o-OCH3) in the presence of [ClRh(CO)2]2 was studied in hexamethylphosphoramide (HMPA). The experimental rate law obtained is ?d[ArHgCl]/dt=k[ArHgCl]2. The kinetic parameters of these reactions have been reported, and the variation, therein, has been explained on the basis of steric effect of substituents. The apparent activation energy E is linearly proportional to pre-exponential factor lnA. A most plausible mechanism has been proposed on the basis of experimental results. © 1993 John Wiley & Sons, Inc.  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