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 93 毫秒
1.
The kinetics of the ring-opening reactions of thiophenolates with geminal bis(acceptor)-substituted cyclopropanes in DMSO at 20 °C was monitored by photometric methods. The determined second-order rate constants of the SN2 reactions followed linear relationships with Mayr nucleophilicity parameters (N/sN) and Brønsted basicities (pKaH) of the thiophenolates as well as with Hammett substituent parameters (σ) for groups attached to the thiophenolates. Phenyl-substituted cyclopropanes reacted by up to a factor of 15 faster than their unsubstituted analogues, in accord with the known activating effect of adjacent π-systems in SN2 reactions. Variation of the electronic properties of substituents at the phenyl groups of the cyclopropanes gave rise to parabolic Hammett relationships. Thus, the inherent SN2 reactivity of electrophilic cyclopropanes is activated by electron-rich π-systems because of the more advanced C1−C2 bond polarization in the transition state. On the other hand, electron-poor π-systems also lower the energetic barriers for the attack of anionic nucleophiles owing to attractive electrostatic interactions.  相似文献   

2.
Hydrolysis reactions of benzyl chlorides and benzenesulfonyl chlorides were theoretically investigated with the density functional theory method, where the water molecules are explicitly considered. For the hydrolysis of benzyl chlorides (para‐Z? C6H4? CH2? Cl), the number of water molecules (n) slightly influences the transition‐state (TS) structure. However, the para‐substituent (Z) of the phenyl group significantly changes the reaction process from the stepwise (SN1) to the concerted (SN2) pathway when it changes from the typical electron‐donating group (EDG) to the typical electron‐withdrawing one (EWG). The EDG stabilizes the carbocation (MeO? C6H4? CH2+), which in turn makes the SN1 mechanism more favorable and vice versa. For the hydrolysis of benzenesulfonyl chlorides (para‐Z? C6H4? SO2? Cl), both the Z group and n influence the TS structure. For the combination of the large n value (n > 9) and EDG, the SN2 mechanism was preferred. Conversely, for the combination of the small n value and EWG, the SN3 one was more favorable. © 2014 Wiley Periodicals, Inc.  相似文献   

3.
To approach more realistic mechanisms for asymmetric aziridine synthesis from guanidinium ylides and aryl aldehydes, reactions were systematically carried out by using a variety of p-substituted benzaldehydes under modified conditions. Two kinds of reaction mechanisms controlled by the nature of the p-substituents of aryl aldehydes is proposed for the two-steps aziridine synthesis composed of a C-C bond formation by nucleophilic addition of guanidinium ylides to aryl aldehydes (step 1) and the fragmentation of intermediate adducts to aziridine products by intramolecular nucleophilic substitution (step 2). A SNi-like mechanism via cationic-like transition state is proposed for step 2 in the asymmetric synthesis using EDG-substituted benzaldehydes, whereas with EWG-substituted benzaldehydes, a SN2-like mechanism is proposed. Hammett analysis, based on the diastereomeric ratio in the aziridine products, is consistent with the proposed rate-determining steps in these two mechanisms. A second Hammett analysis, based on the enantiomeric ratio of the aziridine products, clearly reveals the difference in the susceptibilities to the electronic substituents effect between step 1 and step 2.  相似文献   

4.
The possibility that the similarity of Hammett ρ values for several very different reactions of 2-arylaziridinium salts is due to ground state effects was investigated by MO methods and ESCA. Minimum energy structures of p-nitrophenyl-, phenyl-, and p-methoxyphenyl-substituted ions, however, have the same C2? N bond lengths. This result shows that substituent effects are primarily a transition state phenomenon, even in the three-membered ring heterocycles. In agreement, the ground state does not have significant amino carbocation character based on similar charge distributions of 2-phenyl-N,N-dimethylaziridinium ion and N,N,N-trimethylbenzylammonium ion as well as the same nitrogen ls binding energies by ESCA studies on corresponding salts.  相似文献   

5.
The effects of substituents on the rates of bimolecular substitution of (benzyl chlorides)-chromium tricarbonyl with sodium thiocyanate in acetone are described. Except for the p-methoxy compound, all the complexes react more slowly than the corresponding uncomplexed benzyl chlorides. The decrease in reactivity on complexing is ascribed to the incursion of steric effects. While in the uncomplexed series there is a spectrum of transition states ranging from fully synchronous SN2 to SN1-like, complexation shifts the spectrum towards transition state structures with a well pronounced carbonium-like character. Thus, of the two opposing effects that the tricarbonyl chromium can produce at the α-reaction center, its electron-donating ability prevails over its electron-attracting effect.  相似文献   

6.
A simple potentiometric method for determining the tctraphenylborate ion (C6H5)4B- contents of organic amine tetraphcnylborate salts has been developed. The compounds arc dissolved in 1 : 1 aqueous acetone; and the solution is buffered at a pH of 5 with 3M acetic acid and 3M sodium acetate. The resulting solution is then titrated with 0.06N aqueous silver nitrate, using a silver indicating electrode, a glass reference electrode and a Beckman pH meter to indicate EMF changes. Platinum electrodes in conjunction with the Malmstadt automatic titrator can also be used. Good end-points were obtained for all structure types except the primary aromatic amine borates. Values for (C6H5)4B- found were within 2% of theory for 22 compounds so analyzed.  相似文献   

7.
The gas‐phase ionic SN2 reactions NCS + CH3F and ion pair SN2 reaction LiNCS + CH3F with inversion mechanism were investigated at the level of MP2(full)/6‐311+G**//HF/6‐311+G**. Both of them involve the reactants complex, inversion transition state, and products complex. There are two possible reaction pathways in the ionic SN2 reaction but four reaction pathways in the ion pair SN2 reaction. Our results indicate that the introduction of lithium significantly lower the reaction barrier and make the ion pair displacement reaction more facile. For both ionic and ion pair reaction, methyl thiocyanate is predicted to be the major product, but the latter is more selective. More‐stable methyl isothiocyanate can be prepared by thermal rearrangement of methyl thiocyanate. The theoretical predictions are consistent with the known experimental results. © 2004 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

8.
We report here the first observation of alkali‐metal ion catalysis and inhibition in SNAr reactions. The plot of kobsd versus [alkali‐metal ethoxide] exhibits downward curvature for the reactions of 1‐(4‐nitrophenoxy)‐2,4‐dinitrobenzene with EtOLi, EtONa, and EtOK, but upward curvature for the corresponding reaction with EtOK in the presence of 18‐crown‐6‐ether (18C6). Dissection of kobsd into the second‐order rate constants for the reactions with the dissociated EtO? and the ion‐paired EtOM (i.e., k and kEtOM, respectively) has revealed that the reactivity increases in the order EtOLi<EtONa<EtOK<EtO?<EtOK/18C6. This indicates that the reaction is inhibited by Li+, Na+, and K+ ions but is catalyzed by 18C6 K+ ion. The reactions of 1‐(Y‐substituted‐phenoxy)‐2,4‐dinitrobenzenes have been proposed to proceed through a stepwise mechanism, in which expulsion of the leaving group occurs after the rate‐determining step based on the kinetic result that σo constants exhibit a much better Hammett correlation than σ? constants. Alkali‐metal ion catalysis or inhibition has been discussed in terms of differential stabilization of ground‐state and transition‐state complexes through a qualitative energy profile. A π‐complexed transition‐state structure is proposed to account for the kinetic results.  相似文献   

9.
Radiolysis of aqueous sodium nitrate solution was studied as a function of concentration in the range 10–4M to 1M NaNO3. The radiolytic yield of nitrite was found to be linear with dose and concentration. The effect of transition metal chloride additives on the radiolysis of 0.01M NaNO3 resulted in higher and lower yields of nitrite in the presence of cobalt and nickel chlorides, respectively, than that obtained in the pure nitrate system. The reduction of nitrate to nitrite is totally quenched even at very low concentration of copper chloride in the binary mixture. The results are explained on the basis of oxidizing and reducing properties of transition metal ions.  相似文献   

10.
The decomposition of silver selenide and sulfide to metallic silver and chalcogen containing oxygen compounds by sintering with an equimolar mixture of sodium nitrate and nitrite was examined. It was found that 100% recovery of silver in a metal phase is reached at 5% excess of sodium nitrate and nitrite and a time of the isothermal exposure at 375 °C 1 hour or half an hour at 400°C.  相似文献   

11.
Friedel–Crasfts alkylation reactions of α,β-unsaturated butyric aldehydes with N,N-dimethyl-3-anisidine catalyzed by a (2S,5S)-5-benzyl-2-tert-butyl-3-methylimidazolidin-4-one HCl salt have been carried out at the PCM(CH2Cl2)/B3LYP/6-311++G(d,p)//B3LYP/6-31G(d) level. Three reaction processes have been characterized: (I) the formation of an iminium ion intermediate; (II) the 1,4-iminium addition of the iminium ion; and (III) the hydrolysis of the addition product. Moreover, Path 1-1 is the favorable channel in the formation of the iminium ion. From the point of view of energy, the enantioselectivity is controlled by the carbon–carbon bond formation step that is involved in both the intermediate M4 and the transition state TS4. The highest energy barrier of the reaction is the H2 proton transfer from the O10 atom of a water molecule to the N1 atom of the catalyst in the hydrolysis process, which is 23.4 kcal/mol. The presented calculated results may be helpful in understanding the experimental product distribution for the title reaction, and provide a general model to help explain the mechanisms of similar reactions.  相似文献   

12.
The ρ values of free radical SH2′ reactions have been determined in various solvents. The correlation of Hammett ρ with Taft's π* gives a W value of 0.70. The W value is a measure of susceptibility of the reaction constant to change in solvent polarity. However, the W value is 2.64 in the dissociation reactions of substituted benzoic acids. The free radical reactions are less susceptible to the solvent effect than ionic reactions and this could be rationalized in terms of the partial charge formed in the transition state of free radical reaction is less than that of heterolytic reaction. The ρ values in SH2′ reactions might not reflect truly the partial charge separation at transition state, however, it might be a measure of the susceptibility of the reaction to the electronic effect of the substituents.  相似文献   

13.
The standard potentials of the silver—silver tungstate, silver—silver phosphate and silver—silver arsenate electrodes in four different compositions of water—dioxane and water—urea mixtures at seven different temperatures from 5 to 35°C have been determined from EMF measurements of cells of the type Ag(s), AgCl(s), NaCl(c)//NaxZ(c/x), AgxZ(s), Ag(s), where x is 2 or 3, and Z is WO4, PO4 or AsO4. These values have been used to evaluate the transfer thermodynamic quantities accompanying the transfer of 1 g ion of WO2?4. PO3?4 or AsO3?4 ion from the standard state in water to the standard state in water—dioxane or water—urea mixtures.  相似文献   

14.
This review is devoted to the stereochemistry of nucleophilic substitution reactions at phosphorus. The study of the reactions of phosphoryl group transfer is important for biological and molecular chemistry. The stereochemistry and mechanisms of SN1(P) monomolecular and SN2(P) bimolecular nucleophilic substitution reactions of organophosphorus compounds are discussed. It has been shown that hydrolysis of many natural phosphates proceeds according to the monomolecular SN1(P) mechanism via the formation of metaphosphate intermediate (PO3?). SN2(P) nucleophilic substitution at chiral trivalent or pentavalent phosphorus compounds proceeds via the formation of penta-coordinated transition state or pentacoordinate intermediate.  相似文献   

15.
The anionic SN2 reactions NCO? + CH3X and ion pair SN2 reactions LiNCO + CH3X (X = F, Cl, Br, and I) at saturated carbon with inversion and retention mechanisms were investigated at the level of MP2/6‐311+G(d,p). There are two possible reaction pathways in the anionic SN2 reactions, but eight in the ion pair SN2 reactions. Calculated results suggest that the previously reported T‐shaped isomer of lithium isocyanate does not exist. All the retention pathways are not favorable based on the analysis of transition structures. Two possible competitive reaction pathways proceed via two six‐member ring inversion transition structures. It is found that there are two steps in the most favorable pathway, in which less stable lithium cyanate should be formed through the isomerization of lithium isocyanate and nucleophilic site (N) subsequently attacks methyl halides from the backside. The thermodynamically and kinetically favorable methyl isocyanate is predicted as major product both in the gas phase anionic and the ion pair SN2 reactions. In addition, good correlations between the overall barriers relative to separated reactants, ΔH , with geometrical looseness parameter %L and the heterolytic cleavage energies of the C? X and Li? N (or Li? O) bonds are observed for the anionic and ion pair SN2 reactions. The trend of variation of the overall barriers predicts the leaving ability of X increase in the order: F < Cl < Br < I. The polarized continuum model (PCM) has been used to evaluate the solvent effects on the two inversion pathways with six‐member transition structures for the reactions of LiNCO + CH3X. The calculations in solution indicate that solvent effects will retard the rate of reactions and the predicted product, methyl isocyanate, is same as the one in the gas phase. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  相似文献   

16.
Solvent isotope effects in the ethanolysis of sterically hindered arenesulfonyl chlorides ruled out a proton transfer in the rate‐determining step and agreed with a SN2 mechanism involving at least a second solvent molecule in the transition state (TS). The lack of a secondary kinetic isotope effect in the o‐alkyl groups allows us to disregard the possible contribution of σ–π hyperconjugation. The measured activation parameters are consistent with a SN2 mechanism involving the participation of solvent molecules in the TS, possibly forming a cyclic TS through a chain of solvent molecules.  相似文献   

17.
Syntheses of N,N′-dibutylbenzimidazolylidene silver complexes having chloride, nitrate or cyanide as an anion part through an iodide/anion exchange from N,N′-dibutylbenzimidazolium iodide are described, representing a practical route to benzimidazolylidene silver complexes from readily accessible benzimidazolium iodide. The crystal structures of N,N′-dibutylbenzimidazolylidene silver chloride, bromide, cyanide and nitrate have been determined, showing a close ligand-unsupported Ag-Ag interaction in [(NHC)2]Ag+[AgX2] and a “T” shape geometry about the silver(I) cation in complexes of chloride, bromide and cyanide, but a nearly linear shape in the bis(N,N′-dibutylbenzimidazolylidene) silver complex [ with non-coordinating nitrate anion.  相似文献   

18.
Pseudo acid chlorides derived from levulinic acid ando-benzoyl-benzoic acid, solvolyse in aqueous acetone, aqueous dioxane and aqueous dimethylformamide by aS Nl process. Their reaction pattern is distinct from that of typical normal acid chlorides, viz.,p-benzoylbenzoyl chloride and fluorene-9-one-1-carboxylic acid chloride, which solvolyse by aS N2 pathway. No evidence for tautomerism could be obtained either between the normal and pseudo forms of the acid chlorides or the derived ion pairs.  相似文献   

19.
Alcohols and phenols are efficiently nitrated with thionyl chloride nitrate or thionyl nitrate, even in the presence of an aromatic moiety. While thionyl chloride nitrate is suitable for nitration of primary OH-groups in carbohydrates, thionyl nitrate is reactive enough to react with secondary OH-groups as well. These reagents permit the highly selective nitration of the 5′-,2′5′- and 3′, 5′-OH-groups of ribonucleosides to produce either mono-or diprotected nitro derivatives in high yields. Carbon acids and the enol form of some ketones are efficiently nitrated with trifluoromethanesulfonyl nitrate/potassium tert-butoxide. Lutidine N-oxide (2,6-(CH3)2C5H3N O) was found to have a marked effect on nitration reactions. Similarly, thionyl chloride nitrite and thionyl nitrite exhibit an excellent capacity for nitrosation of the aforementioned substrates.  相似文献   

20.
N‐Heterocyclic carbenes (NHCs) are of great importance and are powerful ligands for transition metals. A new series of sterically hindered benzimidazole‐based NHC ligands (LHX) ( 2a , 2b , 2c , 2d , 2e , 2f ), silver–NHC complexes ( 3a , 3b , 3c , 3d , 3e , 3f ) and palladium–NHC complexes ( 4a , 4b , 4c , 4d , 4e , 4f ) have been synthesized and characterized using appropriate spectroscopic techniques. Studies have focused on the development of a more efficient catalytic system for the Suzuki coupling reaction of aryl chlorides. Catalytic performance of Pd–NHC complexes and in situ prepared Pd(OAc)2/LHX catalysts has been investigated for the Suzuki cross‐coupling reaction under mild reaction conditions in aqueous N,N‐dimethylformamide (DMF). These complexes smoothly catalyzed the Suzuki–Miyaura reactions of electron‐rich and electron‐poor aryl chlorides. Copyright © 2014 John Wiley & Sons, Ltd.  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