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 109 毫秒
1.
Second‐order rate constants for the reactions of acceptor‐substituted phenacyl (PhCO?CH??Acc) and benzyl anions (Ph?CH??Acc) with diarylcarbenium ions and quinone methides (reference electrophiles) have been determined in dimethylsulfoxide (DMSO) solution at 20 °C. By studying the kinetics in the presence of variable concentrations of potassium, sodium and lithium salts (up to 10?2 mol L?1), the influence of ion‐pairing on the reaction rates was examined. As the concentration of K+ did not have any influence on the rate constants at carbanion concentrations in the range of 10?4–10?3 mol L?1, the acquired rate constants could be assigned to the reactivities of the free carbanions. The counter ion effects increase, however, in the series K+<Na+<Li+, and the sensitivity of the carbanion reactivities toward variation of the counter ion strongly depends on the structure of the carbanions. The reactivity parameters N and sN of the free carbanions were derived from the linear plots of log k2 against the electrophilicity parameters E of the reference electrophiles, according to the linear‐free energy relationship log k2(20 °C)=sN(N+E). These reactivity parameters can be used to predict absolute rate constants for the reactions of these carbanions with other electrophiles of known E parameters.  相似文献   

2.
The rates of the hydride abstractions from the 2‐aryl‐1,3‐dimethyl‐benzimidazolines 1a – f by the benzhydrylium tetrafluoroborates 3a – e were determined photometrically by the stopped‐flow method in acetonitrile at 20 °C. The reactions follow second‐order kinetics, and the corresponding rate constants k2 obey the linear free energy relationship log k2(20 °C)= s(N+E), from which the nucleophile‐specific parameters N and s of the 2‐arylbenzimidazolines 1a – c have been derived. With nucleophilicity parameters N around 10, they are among the most reactive neutral C? H hydride donors which have so far been parameterized. The poor correlation between the rates of the hydride transfer reactions and the corresponding hydricities (ΔH0) indicates variable intrinsic barriers.  相似文献   

3.
TiCl4‐induced Baylis–Hillman reactions of α,β‐unsaturated carbonyl compounds with aldehydes yield the (Z)‐2‐(chloromethyl)vinyl carbonyl compounds 5 , which react with 1,4‐diazabicyclo[2.2.2]octane (DABCO), quinuclidine, and pyridines to give the allylammonium ions 6 . Their combination with less than one equivalent of the potassium salts of stabilized carbanions (e.g. malonate) yields methylene derivatives 8 under kinetically controlled conditions (SN2’ reactions). When more than one equivalent of the carbanions is used, a second SN2’ reaction converts 8 into their thermodynamically more stable allyl isomers 9 . The second‐order rate constants for the reactions of 6 with carbanions have been determined photometrically in DMSO. With these rate constants and the previously reported nucleophile‐specific parameters N and s for the stabilized carbanions, the correlation log k (20 °C)=s(N + E) allowed us to calculate the electrophilicity parameters E for the allylammonium ions 6 (?19<E <?18). The kinetic data indicate the SN2’ reactions to proceed via an addition–elimination mechanism with a rate‐determining addition step.  相似文献   

4.
Thioether 4‐[(1′E,3′E)‐4′‐phenylsulfanyl‐1,3′‐butadienyl]pyridine 8 and sulfone 4‐(4′‐phenylsulfonyl‐1′,3′‐butadienyl)pyridine 14 were prepared by reaction of the carbanions derived from allylic thioether or allylic sulfone with isonicotinaldehyde. The reaction with the sulfonyl carbanion occurred at the α position and on heating the alcolate gave the dienic sulfone 14 . The corresponding pyridinium iodide 10 and 15 were prepared by reaction with methyl iodide, respectively, on pyridine derivates 8 and 14 . The dienic pyridinium thioether 10 showed a long wavelength absorption band centered at 420 nm. The reaction of dienic pyridinium sulfone 15 with thiophenol gave the dienic pyridinium thioether 10 by a nucleophilic vinylic substitution. The reaction of sulfone 15 with glutathione was of second order and the rate constant was 8.5 M?1s?1 at 30°C and pH 7, about 500 times smaller than the rate constant observed with (E)‐1‐methyl‐4‐(2‐methylsulfonyl‐1‐ethenyl)pyridinium iodide 1 . The dienic pyridinium thioether 10 was a negative solvatochrome.  相似文献   

5.
Kinetics of the reactions of 3,5‐dinitrothiophene 1 and 3‐cyano‐5‐nitrothiophene 2 with a series of parasubstituted phenoxide anions 3a–c have been investigated in aqueous solution at 20°C. Two unsubstituted electrophilic centers (C(2) and C(4)) of the two thiophenes have been identified. The Fukui functions correctly predict the C(2) and C(4) atoms as the most electrophilic centers of these electron‐deficient thiophenes 1 and 2 . Analysis of the experimental data in terms of Brønsted relationships reveals that the reaction mechanism likely involves a single‐electron transfer (SET) process. The excellent correlations upon plotting the rate constants versus the oxidation potentials Eo values is an additional evidence that reactions between thiophenes and phenoxide anions are proceeding through an initial electron transfer. It is of particular interest to note that the systems studied in this paper provide a rare example of a SET mechanism in σ‐complexation reactions. According to the free energy relationship log k = s(N + E) (Angew. Chem., Int. Ed. Engl., 1994, 33, 938–957), the electrophilicity parameters E of the C‐4 and C‐2 positions of the thiophenes have been determined and compared with the reactivities of other ambident electrophiles. On the other hand, the second‐order rate constants for the reactions of these thiophenes with the hydroxide ion has been measured in water and 50% water–50% acetonitrile and found to agree with those calculated theoretically using Mayr's equation from the E values determined in this work and from the previously published N and s parameters of OH.  相似文献   

6.
Diethylbis(2,2′‐bipyridine)Fe/MAO is an extremely active catalyst for the polymerization of 1,3‐dienes. Polymers with a 1,2 or 3,4 structure are formed from butadiene, isoprene, (E)‐1,3‐pentadiene and 3‐methyl‐1,3‐pentadiene, while cis‐1,4 polymers are derived from 2,3‐dimethyl‐1,3‐butadiene. The 1,2 (3,4) polymers obtained at 25°C are amorphous, while those obtained below 0°C are crystalline, as was determined by means of X‐ray diffraction. Mechanistic implications of the results are briefly discussed.  相似文献   

7.
The late transition metal catalyzed rearrangement of propargyl acetates offers an interesting platform for the development of synthetically useful transformations. We have recently shown that gold complexes can catalyze a highly selective tandem 1,2‐/1,2‐bis‐acetoxy migration in 1,4‐bis‐propargyl acetates to form 2,3‐bis‐acetoxy‐1,3‐dienes. In this way, (1Z,3Z)‐ or (1Z,3E)‐ and (1E,3Z)‐1,3‐dienes could be obtained in a stereocontrolled manner depending on the electronic and steric features of the ancillary ligand bound to gold and the substituents at the propargylic positions. In this work, we report an experimental study on the scope of this transformation, plus a detailed theoretical examination of the reaction mechanism, which has revealed the key features responsible for the reaction stereoselectivity. Synthetic applications towards the one‐pot synthesis of quinoxaline heterocycles and tandem Diels–Alder processes have also been devised.  相似文献   

8.
The reaction of dialkyl acetylenedicarboxylates with arylidenemalononitriles in the presence of KSCN in MeCN led to a mixture of dialkyl (3E)‐4‐aryl‐3‐(arylideneamino)‐5,5‐dicyanocyclopenta‐1,3‐diene‐1,2‐dicarboxylates and dialkyl 4‐aryl‐5‐cyanothiophene‐2,3‐dicarboxylates. When these reactions were performed in the presence of KCN, only the functionalized 5,5‐dicyanocyclopenta‐1,3‐dienes were obtained.  相似文献   

9.
(E)‐2‐[2‐(1‐Substituted ethylidene)hydrazinyl]‐5‐oxo‐9b‐hydroxy‐5,9b‐dihydroindeno[1,2‐d][1,3]‐thiazine‐4‐carbonitriles and (E)‐5‐oxo‐[(E)‐(1‐substituted ethylidene)hydrazinyl]‐2,5‐dihydroindeno[1,2‐d][1,3]thiazine‐4‐carbonitriles have been obtained from the reaction of 2‐(substituted ethylidene)hydrazinecarbothioamides with 2‐(1,3‐dioxo‐2,3‐dihydro‐1H‐inden‐2‐ylidene)propanedinitrile ( 1 ) in ethyl acetate solution. However, (Z)‐6′‐amino‐1,3‐dioxo‐3′‐substituted‐2′‐[(E)‐(1‐phenylethylidene)hydrazono]‐1,2′,3,3′‐tetrahydrospiro(indene‐2,4′‐[1,3]thiazine)‐5′‐carbonitriles were observed during the reaction of N‐substituted‐2‐(1‐phenylethylidene)hydrazinecarbothioamides with ( 1 ). The structure assignment of products has been confirmed on the basis of 1H‐, 13C‐NMR, and mass spectrometry, as well as theoretical calculations.  相似文献   

10.
N,N′‐Bis(difuroxano[3,4‐b:3′,4′‐d]phenyl)oxalic amide was synthesized via acylation, nitration, azidation, and pyrolysis‐denitrogenation from the starting materials of oxalyl chloride and 3,5‐dichloroaniline, under mild reaction conditions, with the yields of 81.0%, 82.0%, 86.0% and 81.7% respectively. The title compound and its precursors were characterized by 1H NMR, IR, MS, and elemental analysis. The title compound has a density of 1.92 g·cm?3 by a suspension method, a standard formation enthalpy of 979 kJ·mol?1 calculated by Gaussian programs, a detonation velocity of 8.17 km·s?1, and a detonation pressure of 31 GPa obtained by Kamlet Equation. The thermal decomposition reactions of the title compound at different heating rates were tested by differential scanning calorimetry (DSC). The kinetics parameters of the pyrolysis of the compound were calculated by Kissinger's method. The values of apparent activation energy (Ea) and pre‐exponential constant (A) were 226.7 kJ·mol?1 and 1023.17 s?1 respectively. It was presupposed that N,N′‐bis(difuroxano[3,4‐b:3′,4′‐d]phenyl)oxalic amide would be a promising high energetic explosive with low sensitivity.  相似文献   

11.
When treated with LiNiPr2 (LDA) at ?78°, 1‐[(methylsulfanyl)methyl]‐2‐[(1Z,3E)‐4‐phenylbuta‐1,3‐dien‐1‐yl]benzene easily cyclized to form benzocycloheptenyl anion, which successively underwent intramolecular nucleophilic substitution to give a cyclopropanaphthalene. Similar LDA‐mediated cyclization also occurred for 4‐phenyl‐ or 4‐methyl‐substituted 1‐[2‐(methoxymethyl)phenyl]buta‐1,3‐dienes to furnish the corresponding benzocycloheptenes and cyclopropanaphthalenes. A 4‐tert‐butyl analog also underwent LDA‐mediated cyclization to give a benzocycloheptene, but not a cyclopropanaphthalene.  相似文献   

12.
Kinetic studies for the azo‐coupling reactions of 3‐ethoxythiophene 1 with a series of 4‐X‐substituted diazonium cations 2a‐e (X = OCH3, CH3, H, Cl, and NO2) have been investigated in acetonitrile at 20°C. The second‐order rate constants have been employed to determine the nucleophilicity parameters N and s of the thiophene 1 according the Mayr equation. Thus, the nucleophile‐specific parameters N and s of thiophene 1 have been derived and compared with the reactivities of other C‐nucleophiles in acetonitrile (pyrroles, furan, indoles, etc.). The Yukawa–Tsuno plot resulted in an excellent correlation (R2 = 0.9980) with an r value of 0.89, suggesting that the nonlinear Hammett plot observed in the present work is due to resonance demand of the π–electron donor substituent of on the –N2+ moiety. Importantly, using the concept of global electrophilicity (ω) proposed by Parr, we successfully predict the electrophilicity parameters E of seven substituted diazonium cations whose experimental data are available.  相似文献   

13.
Reaction of O,O′‐diisopropylthiophosphoric acid isothiocyanate (iPrO)2P(S)NCS with 1,10‐diaza‐18‐crown‐6, 1,7‐diaza‐18‐crown‐6, or 1,7‐diaza‐15‐crown‐5 leads to the N‐thiophosphorylated bis‐thioureas N,N′‐bis[C(S)NHP(S)(OiPr)2]‐1,10‐diaza‐18‐crown‐6 ( H2LI ), N,N′‐bis[C(S)NHP(S)(OiPr)2]‐1,7‐diaza‐18‐crown‐6 ( H2LII ) and N,N′‐bis[C(S)NHP(S)(OiPr)2]‐1,7‐diaza‐15‐crown‐5 ( H2LIII ). Reaction of the potassium salts of H2LI–III with a mixture of CuI and 2,2′‐bipyridine ( bpy ) or 1,10‐phenanthroline ( phen ) in aqueous EtOH/CH2Cl2 leads to the dinuclear complexes [Cu2(bpy)2LI–III] and [Cu2(phen)2LI–III] . The structures of these compounds were investigated by 1H, 31P{1H} NMR spectroscopy, and elemental analysis. The crystal structures of H2LI and [Cu2(phen)2LI] were determined by single‐crystal X‐ray diffraction. Extraction capacities of the obtained compounds in comparison to the related compounds 1,10‐diaza‐18‐crown‐6, N,N′‐bis[C(=CMe2)CH2P(O)(OiPr)2]‐1,10‐diaza‐18‐crown‐6, N,N′‐bis[C(S)NHP(O)(OiPr)2]‐1,10‐diaza‐18‐crown‐6 towards the picrate salts LiPic, NaPic, KPic. and NH4Pic were also studied.  相似文献   

14.
1,3‐Dipentafluorophenyl‐2,2,2,4,4,4‐hexazido‐1,3‐diaza‐2,4‐diphosphetidine ( 1 ) was synthesized by the reaction of [(C6F5)NPCl3]2 with trimethylsilyl azide in CH2Cl2 and characterized by multinuclear NMR and vibrational spectroscopy. The molecular structure of the compound was determined by single‐crystal X‐ray structure analysis. [(C6F5)NP(N3)3]2 crystallizes in the monoclinic space group P21/n with a = 9.6414(2), b = 7.4170(1) and c = 15.9447(4) Å, β = 94.4374(9)°, with 2 formula units per unit cell. The bond situation in [(C6F5)NP(N3)3]2 has been studied on the basis of NBO analysis. The antisymmetric stretching vibration of the azide groups is discussed. The structural diversity of 1 and 1,3‐diphenyl‐2,2,2,4,4,4‐hexazido‐1,3‐diaza‐2,4‐diphosphetidine in solution and in the solid state depending on the aryl substituent at the nitrogen atom is discussed.  相似文献   

15.
The title triene, C18H10F6, was prepared via the Pd0 coupling reaction of (E)‐(1,2‐di­fluoro‐1,2‐ethenediyl)­bis­(tri­butyl­stan­nane) with (Z)‐β‐iodo‐α,β‐di­fluoro­styrene in N,N′‐dimethylformamide/tetrahydrofuran. The crystal structure shows the product to be the 1E,3E,5E isomer. Due to steric interactions between F atoms, the double bonds are not coplanar. The planes defined by the two terminal double bonds are almost perpendicular.  相似文献   

16.
A kinetic study is reported for reactions of 2‐methoxy‐3‐X‐5‐nitrothiophenes 1a–d (X = SO2CH3, CO2CH3, CONH2, H) with piperidine in different solvents at 20°C. It is shown that the reactions take place through a SNAr mechanism with the initial nucleophilic addition step being rate limiting. The satisfactory Hammett correlations (log k1 vs. σ) obtained in the present system confirms that a 3‐X substituent exerts an effect on the 2‐position of the same type as that exerted from the 5‐position. The second‐order rate constants associated with these reactions are employed to determine the electrophilicity parameters E of the thiophenes 1a–d according to the relationship log k (20°C) = s(E + N) (Angew. Chem., Int. Ed. Engl. 1994, 33, 938–957). The E values of 1a–d are found to cover a range from ?21.33 to ?17.18, going from 1d , the least reactive, to 1a , the most reactive thiophene. Interestingly, a linear correlation (r2 = 0.9910) between the electrophilicity parameters E determined in this work and the Hammett's σ constants values has been observed and discussed. On the other hand, we have found that the reported rate constants of some thiophenes 1 complexation by the methoxide ion in methanol are 3.5–73.5 times higher than predicted by Mayr's approach.  相似文献   

17.
Kinetics of the reactions of bissulfonyl ethylenes with various carbanions, a sulfur ylide, and siloxyalkenes have been investigated photometrically at 20 °C. The second‐order rate constants have been combined with the known nucleophile‐ specific parameters N and sN for the nucleophiles to calculate the empirical electrophilicity parameters E of bissulfonyl ethylenes according to the linear free energy relationship log k(20 °C)=sN(N+E). Structure‐reactivity relationships are discussed, and it is shown that the electrophilicity parameters E derived in this work can be employed to define the synthetic potential of bissulfonyl ethylenes as Michael acceptors.  相似文献   

18.
A one‐pot synthesis of pyrrolo[1,2‐a]quinolin‐1‐ones has been developed from the reactions of 5‐hydroxy‐1‐arylpyrrolidin‐2‐ones with 1,3‐dicarbonyl compounds under the promotion of H3PO4/P2O5 or HOAc/H2SO4. The pyrrolo[1,2‐a]quinolin‐1‐ones are formed by two‐step reactions, that is, the coupling of N‐acyliminium ion intermediates produced from 5‐hydroxy‐1‐arylpyrrolidin‐2‐ones with 1,3‐dicarbonyls and subsequent Friedel–Crafts reactions of the resulting ketone with the aryl ring.  相似文献   

19.
In the title compound, diaqua­bis(1,4‐di‐4‐pyrid­yl‐2,3‐diaza‐1,3‐butadiene)dimethanolzinc(II) bis­(perchlorate) 1,4‐di‐4‐pyrid­yl‐2,3‐diaza‐1,3‐butadiene methanol 1.72‐solvate 1.28‐hydrate, [Zn(C12H10N4)2(CH4O)2(H2O)2](ClO4)2·C12H10N4·1.72CH4O·1.28H2O, determined at ca 110 K, the Zn cation and the extended dipyridyl ligand both lie across inversion centres in space group P. The structure consists of a network arrangement of the constituent species stabilized by a combination of coordination, hydrogen bonding and π–π forces. Uncoordinated methanol and water solvent mol­ecules occupy the otherwise void spaces within and between the networks.  相似文献   

20.
A new compound, 2‐(dinitromethylene)‐1,3‐diazacyclopentane (DNDZ), was prepared by the reaction of 1,1‐diamino‐2,2‐dinitroethylene (FOX‐7) with 1,2‐diaminoethane in N‐methylpyrrolidone (NMP). Thermal decomposition of DNDZ was studied under non‐isothermal conditions by DSC, TG/DTG methods, and the enthalpy, apparent activation energy and pre‐exponential factor of the exothermic decomposition reaction were obtained as 317.13 kJ·mol?1, 269.7 kJ·mol?1 and 1024.51 s?1, respectively. The critical temperature of thermal explosion was 261.04°C. Specific heat capacity of DNDZ was determined with a micro‐DSC method and a theoretical calculation method, and the molar heat capacity was 205.41 J·mol?1·K?1 at 298.15 K. Adiabatic time‐to‐explosion was calculated to be a certain value between 263–289 s. DNDZ has higher thermal stability than FOX‐7.  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