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 10 毫秒
1.
A ligand‐controlled system that enables regioselective trifluoromethylcyanation of 1,3‐enynes has been identified, which provides access to a variety of CF3‐containing tri‐ and tetrasubstituted allenyl nitriles. We disclose that the involved propargylic radicals can be selectively trapped by (Box)CuII cyanide, while the tautomerized allenyl radicals are trapped by (phen)CuII cyanide (Box= bisoxazoline, phen=phenanthroline). In addition, the reaction features broad substrate scope and excellent functional group compatibility. Moreover, this protocol represents a novel regioselectivity‐tunable functionalization of 1,3‐enynes via radicals, which we believe will have great implications for the development of catalytic systems for selectivity control in radical and organometallic chemistry.  相似文献   

2.
A combination of microcalorimetry, the rotating sector method, and ESR at 323 K in the environment of 10 solvents of different polarities was used to measure rate constants of addition of hydroperoxide radicals () to π bonds of trans‐1,2‐diphenylethylene and trans,trans‐1,4‐diphenylbutadiene‐1,3 (k2) and disproportionation rate constants of these radicals (k3). With increasing dielectric constant of the medium, k2 values increase from 69 to 410 M−1 · s−1, and k3 values almost do not change and are in the range of (1.0 ± 0.2) × 108 M−1 · s−1. A linear dependence of logarithm values of rate constants from the dielectric constant of the medium in the coordinates of the Kirkwood–Onsager equation was found that allows to make a conclusion about the effect of nonspecific solvation in the studied systems. The quantum‐chemical analysis (NWChem, DFT B3LYP/6‐311G**) of the detailed mechanism for addition shows that the influence of the medium polarity reflects the superposition of the effects of nonspecific and specific solvation. The scale of the polar effect will depend on how different solvation energies of the transition and the initial reaction complexes. If a value of the solvation energy of the transition complex is larger than the solvation energy of the initial reaction complex, then the reaction rate should increase with an increase of the solvent's polarity and decrease otherwise.  相似文献   

3.
We provide a seminal example of the utility of the 1,2‐azaborine motif as a 4C+1N+1B synthon in organic synthesis. Specifically, conditions for the practically scalable photoisomerization of 1,2‐azaborine in a flow reactor are reported that furnish aminoborylated cyclobutane derivatives. The C?B bonds could also be functionalized to furnish a diverse set of highly substituted cyclobutanes.  相似文献   

4.
5.
6.
In this study density functional theory (DFT) calculations at B3LYP/6-31G(d), B3LYP/6-31+G(d) and B3LYP/6-311+G(2df,2p) levels for geometry optimization and total energy calculation were applied for investigation of the important energy-minimum conformations and transition-state of 1,2-, 1,3-, and 1,4-dithiepanes. Moreover, ab initio calculations at HF/6-31G(d) level of theory for geometry optimization and MP2/6-311G(d)//HF/ 6-31G(d) level for a single-point total energy calculation were reported for different conformers. The obtained results reveal that, the twist-chair conformer is a global minimum for all of these compounds. Also, two local minimum were found in each case, which are twisted-chair and twisted-boat conformers. The boat and chair geometries are transition states. The minimum energy conformation of 1,2-dithiepane is more stable than the lowest energy forms of 1,3-dithiepane and 1,4-dithiepane. Furthermore, the anomeric effect was investigated for 1,3-dithiepane by the natural bond orbital method. The computational results of this study shows that all conformers of 1,3-dithiepane have a hypercojugation system. Finally, the 13C NMR chemical shifts for the conformers of 1,4-dithiepane were calculated, which have good correlation with their experimental values.  相似文献   

7.
This work describes the synthesis of π‐conjugated polymers possessing arylene and 1,3‐butadiene alternating units in the main chain by the reaction of α,β‐unsaturated ester/nitrile containing γ‐H with aromatic/heteroaromatic aldehyde compound. By using 4‐(4‐formylphenyl)‐2‐butylene acid ethyl ester as a model monomer, the different polymerization conditions, including catalyst, catalyst amount, and solvent, are optimized. The polymerization of 4‐(4‐formylphenyl)‐2‐butylene acid ethyl ester is carried out by refluxing in ethanol for 72 h with 1,8‐diazabicyclo[5.4.0]undec‐7‐ene (DBU) as a catalyst to give a 1,3‐butadiene‐containing π‐conjugated polymer, poly(phenylene‐1,3‐butadiene), in 84.3% yield with and / (PDI) estimated as 6172 and 1.65, respectively. Based on this new methodology, a series of π‐conjugated polymers containing 1,3‐butadiene units with different substituents are obtained in high yields. A possible mechanism is proposed for the polymerization through a six‐membered ring transition state and then a 1,5‐H shift intermediate.

  相似文献   


8.
The conformational analysis of cycloheptane (1), oxacycloheptane (2), 1,2‐dioxacycloheptane (3), 1,3‐dioxacycloheptane (4), and 1,4‐dioxacycloheptane (5) has been carried out using B3LYP, CCD, CCSD, and QCISD with the 6‐311+G(d,p) and cc‐pVDZ basis sets. The twist chair conformers are predicted to be lower in energy than their corresponding boat and chair conformations. All levels of theory predict (4) to be lower in energy than (3) and (5). CCSD predicts remarkably similar activation barriers for the conformational interconversion of the twist chair conformers to their corresponding boat conformers. Small barriers to pseudorotation are also predicted. © 2007 Wiley Periodicals, Inc. Int J Quantum Chem, 2008  相似文献   

9.
Polyesters containing 1,3‐cyclobutylene and 1,4‐cyclohexylene linkages in the main chain are investigated using conformational energy calculations. Rotational isomeric state (RIS) models are developed for poly(1,4‐cyclohexylenedimethylene terephthalate) (PCT), poly(1,4‐cyclohexylenedimethylene 2,5‐dimethylterephthalate) (DMPCT), poly(1,4‐cyclohexylenedimethylene 1,4‐cyclohexylenedicarboxylate) (PCC), and poly(2,2,4,4‐tetramethyl‐1,3‐cyclobutylene terephthalate) (CBDO). In DMPCT, the ester linkage prefers skewed conformations with respect to the phenyl group to relieve the steric strain caused by the methyl groups. The methyl groups on the cyclobutanediol moiety in CBDO restrict the rotational freedom about the oxycyclobutylene linkage. The unperturbed dimensions as described by characteristic ratio and persistence length are calculated for the trans and cis configurations of these polyesters. CBDO shows highly extended chain conformations among these polyesters indicating relative chain rigidity of the backbone. For DMPCT and PCC, in their trans configuration, the chain dimensions decrease with an increase in temperature while for their cis configurations, the chain dimensions increase with temperature, arising from basic differences in the fragment structures that control the competition of the relative populations as affected by temperature. Temperature has negligible influence on the unperturbed dimensions of both isomeric linkages of CBDO, while this is true for the trans configuration of PCT. The study shows that induction of cyclobutylene groups in the main chain results in a greater rigidity for homopolyesters than for chains with cyclohexylene groups.

Structure of the repeat units of polyesters in the planar trans configuration and the schematic of the RIS models with definition of geometrical parameters.  相似文献   


10.
A site‐selective hydroxylative dearomatization of 2‐substituted phenols to either 1,2‐benzoquinols or their cyclodimers, catalyzed by 4,5‐dimethyl‐2‐iodoxybenzenesulfonic acid with Oxone, has been developed. Natural products such as biscarvacrol and lacinilene C methyl ether could be synthesized efficiently under mild reaction conditions. Furthermore, both the reaction rate and site selectivity could be further improved by the introduction of a trialkylsilylmethyl substituent at the 2‐position of phenols. The corresponding 1,2‐quinols could be transformed into various useful structural motifs by [4+2] cycloaddition cascade reactions.  相似文献   

11.
The kinetics of the reactions of 1,2‐diaza‐1,3‐dienes 1 with acceptor‐substituted carbanions 2 have been studied at 20 °C. The reactions follow a second‐order rate law, and can be described by the linear free energy relationship log k(20 °C)=s(N+E) [Eq. (1)]. With Equation (1) and the known nucleophile‐specific parameters N and s for the carbanions, the electrophilicity parameters E of the 1,2‐diaza‐1,3‐dienes 1 were determined. With E parameters in the range of ?13.3 to ?15.4, the electrophilic reactivities of 1 a–d are comparable to those of benzylidenemalononitriles, 2‐benzylideneindan‐1,3‐diones, and benzylidenebarbituric acids. The experimental second‐order rate constants for the reactions of 1 a – d with amines 3 and triarylphosphines 4 agreed with those calculated from E, N, and s, indicating the applicability of the linear free energy relationship [Eq. (1)] for predicting potential nucleophilic reaction partners of 1,2‐diaza‐1,3‐dienes 1 . Enamines 5 react up to 102 to 103 times faster with compounds 1 than predicted by Equation (1), indicating a change of mechanism, which becomes obvious in the reactions of 1 with enol ethers.  相似文献   

12.
Irradiation of unsymmetrically substituted 1,4‐dihydropyridines ( 1b ‐ 1j ) while bubbling oxygen or argon through the solution resulted in aromatization to the corresponding pyridine derivatives ( 3b ‐ 3j ). Compound 1a with 2‐nitrophenyl substituent in 4‐position undergoes elimination of water upon irradiation under both oxygen and argon atmospheres and formation of 3a with 2‐nitrosophenyl substituent. On the other hand, irradiation of the compounds 1e , 1k and 1l with 4‐hydroxy‐3‐methoxyphenyl, 5‐methyl‐2‐furyl and 2‐furyl substituent in this position, respectively, resulted in the expulsion of these substituents and formation of a pyridine derivative unsubstituted in position 4, namely compound 2 . Chloroform as a solvent causes the photo‐oxidation of these compounds by an electron transfer mechanism which is supported by the formation of dichloromethane according to GC analysis and presence of acid (HCl) in the solution after irradiation.  相似文献   

13.
The continued use of fossil fuels as primary sources of energy in industry and other applications stands the test of time, due to their availability and relatively lower cost than alternative sources of energy. In view of this perspective, obtaining an advanced bulk carbon dioxide (CO2) capture medium becomes an urgent necessity so as to mitigate their effect, especially in global warming, as the use of fossil fuels produces a high rate of CO2. In this work, the mechanism and kinetics of CO2 capture using aqueous piperazine (PZ) as an activator to 2‐amino‐2‐methyl‐1,3‐propanediol (AMPD) were investigated. The termolecular mechanism was used to model the kinetics of the system. Reaction kinetics of the single pure amines was first obtained. The reaction rate constant, the k value of AMPD, was 77.2 m3/kmol·s, with a reaction order, n, of 1.25 at 298 K. while that of PZ was equal to 11,059 m3/kmol·s and n as 1.49 at 298 K. Blending of 0.05 kmol/m3 of PZ with 0.5 kmol/m3 of AMPD gave a rate constant, k, value of 23,319 m3/kmol·s and n equal to 1.23 at 298 K. The result obtained for the blended system is more than twice the value of the summation of the corresponding pure amines; in addition, it is comparably higher than the rate constant of monoethanolamine (MEA) in use as a commercial solvent for CO2 capture. Therefore, an aqueous blend of PZ with AMPD deserves more comprehensive study as a solvent for commercial CO2 capture. AMPD like other sterically hindered amines absorbs CO2 in an equimolar ratio that is significantly higher than that of MEA. PZ serves as a promoter in the amine mixture and is required in a very small proportion.  相似文献   

14.
15.
An efficient procedure for the synthesis of 7‐(aryl)‐8‐nitro‐2,3,6,7‐tetrahydroimidazo[1,2‐a]pyridinones, 8‐(aryl)‐9‐nitro‐3,4,7,8‐tetrahydropyridone[1,2‐a]pyrimidines and 9‐(aryl)‐10‐nitro‐2,3,4,5,8,9‐hexahydropyridone[1,2‐a]diazepine via one‐pot three component reaction of diamine, nitroketene dithioacetal (1,1‐bis(methylsulfanyl)‐2‐nitroethene), and coumarine‐3‐ carboxylic acid derivatives in EtOH under reflux conditions is reported. The advantages of this procedure are simplicity, easy purification, good yields, and catalyst‐free conditions. All products were confirmed by 1H‐ and 13C‐NMR, IR, MS, and X‐ray crystal structure analyses.  相似文献   

16.
Bismuth nitrate catalyzed condensation reactions of indoline with 1,2‐ and 1,3‐diketones were investigated and were reported to proceed via different reaction pathways with the involvement of one or two of the carbonyl groups. While the reaction of indoline with cyclohexane‐1,3‐dione ( 4 ) gave solely condensation product, the reaction between the acetylacetone ( 5 ) and indoline provided N‐acetyl indoline as single products on retro‐aldol process. In contrast to 1,3‐diketones, the reaction with benzil ( 17 ) was performed under difficult conditions and proceeded to give secondary products.  相似文献   

17.
We report the single crystal structures of 1,4‐bis­(triisopropyl­silyl)buta‐1,3‐diyne, C22H42Si2, and 1,4‐bis­(biphenyl‐4‐yl)buta‐1,3‐diyne, C28H18, the packing in both of which illustrates the versatility of weak C—H⋯π supra­molecular inter­actions in dictating the overall solid‐state structures.  相似文献   

18.
19.
20.
The kinetics of the oxidation of substituted 4-oxoacids by N-bromosaccharin (NBSac) has been studied in aqueous acetic acid medium at 30 °C. The reactions follow first-order kinetics in the 4-oxoacids, NBSA and H+. Variation in the ionic strength has no effect on the reaction rate. The order of reactivity among the studied 4-oxoacids is: 4-methoxy > 4-methyl > 4-phenyl > 4-H > 4-Cl > 4-Br > 3-NO2. The effect of changes in the electronic nature of the substrate revealed that there is a development of positive charge in the transition state. The activation parameters were computed from an Arrhenius plot. Based on the kinetic results, a suitable mechanism has been proposed. The mechanism involves the attack of the oxidizing species hypobromous acidium ion, (H2O+Br).  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