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 8 毫秒
1.
The basis dehydrobromination of p-substituted 2-phenylethyl bromides promoted by hydroxide ion has been studied in micelles of N-hexadecyl-N,N,N-trimethylammonium nitrate (CTANO3). The pseudophase ion exchange model can be adapted to explain the kinetic results at low hydroxide ion concentration, i.e., less than 0.075 M of NaOH. At high hydroxide ion concentration in solution the applied model does not describe the kinetic results unless the empirical variations are applied for the mol ratio of reactive ion to micellar head group (β) and the binding constant of the organic reagent to the micelle (Ks).  相似文献   

2.
Pyridine N-oxide does not react directly with α-bromoiso-butyric acid or with its ethyl ester, but the resulting adducts can be obtained easily as crystallized nitrates 1 and 2 if the reaction is performed in the presence of silver nitrate in acetonitrile solution. Although salt 2 bearing an ester group, does not undergo solvolysis of its O-Cα bond, and cannot be decomposed according to the classical mode A, the kinetic study of its ring opening by hydroxide ions cannot be performed free of any side reaction; hydrolysis of the ester group, followed by opening of the resulting betaine accounts for an observed order of reaction greater than 2 with respect to hydroxide ion. This ring opening competes with a decomposition to carbon dioxide, acetone and pyridine, which is carboxylog of the mode A and which occurs when salt 1 is treated by hydroxide ions.  相似文献   

3.
Reactions of 2,2-dimethylaziridine with benzohydroximoyl chlorides [ArC(Cl)?NOH] give aziridinylbenzaldoximes 1 . It has been found that the aziridine ring in these compounds undergoes ring opening in hydrogen chloride-dioxane solution to give (Z)-N-hydroxy-N′-(2-chloro-2-methylpropyl)benzenecarboximidamides [ArC(NHCH2CR1R2Cl)?NOH, 4 ]. Treatment of 1 with hydrochloric acid followed by neutralization with aqueous sodium hydroxide gave 6,6-dimethyl-3-aryl-1,2,4-oxadiazines 2. Reaction of 4 with sodium hydride in dioxane gave 5-isopropyl-3-aryl-4,5-dihydro-1,2,4-oxadiazoles 5. Reaction of the 4,5-dihydro-1,2,4-oxadiazoles 5 with N-chlorosuccinimide gave the heteroaromatic 1,2,4-oxadiazoles 6 . It is suggested that reactions of 4 with sodium hydride in dioxane solution involve the conjugate base of 4 which undergoes a 1,2-hydride shift that is concerted with loss of chloride ion. In aqueous sodium hydroxide solution it is suggested that the conjugate base of 4 undergoes ionization of the chlorine atom followed by nucleophilic attack by the oximate anion.  相似文献   

4.
It has been shown previously that the reaction of diazomethane with 5‐benzylidene‐3‐phenylrhodanine ( 1 ) in THF at ?20° occurs at the exocyclic C?C bond via cyclopropanation to give 3a and methylation to yield 4 , respectively, whereas the corresponding reaction with phenyldiazomethane in toluene at 0° leads to the cyclopropane derivative 3b exclusively. Surprisingly, under similar conditions, no reaction was observed between 1 and diphenyldiazomethane, but the 2‐diphenylmethylidene derivative 5 was formed in boiling toluene. In the present study, these results have been rationalized by calculations at the DFT B3LYP/6‐31G(d) level using PCM solvent model. In the case of diazomethane, the formation of 3a occurs via initial Michael addition, whereas 4 is formed via [3+2] cycloaddition followed by N2 elimination and H‐migration. The preferred pathway of the reaction of 1 with phenyldiazomethane is a [3+2] cycloaddition, subsequent N2 elimination and ring closure of an intermediate zwitterion to give 3b . Finally, the calculations show that the energetically most favorable reaction of 1 with diphenyldiazomethane is the initial formation of diphenylcarbene, which adds to the S‐atom to give a thiocarbonyl ylide, followed by 1,3‐dipolar electrocyclization and S‐elimination.  相似文献   

5.
The fragmentation patterns of N-phosphoamino acids in negative-ion fast atom bombardment mass spectrometry (FABMS) showed different characteristics to those in positive-ion FABMS. Six typical N-diisopropyloxyphorphorylamino acids all had intense [M ? 1]? peaks, and they underwent similar fragmentation pathways. In general, the elimination of one alkene molecule followed by the loss of one molecule of alcohol occurred. They also favoured an N → O rearrangement reaction, followed by fragmentation to (RO)2 PO2? and (RO) (HO)PO2?.  相似文献   

6.
Phenols containing in the para-position of the benzene ring substituents of various electronic character add to 1,1-difluorodichloroethene in acetone in the presence of potassium hydroxide. A similar reaction with 1,2-di(fluorochloro)ethene occurs only in DMF or N,N-dimethylacetamide and is followed by hydrogen chloride elimination. Phenols with electron-donor substituents add to trifluorochloroethene in acetone in the presence of potassium hydroxide, the reaction of phenols with electron-acceptor substituents requires DMF as solvent.  相似文献   

7.
Electron impact positive ion spectra of ten substituted or annelated benzofurazan-1-oxides are reported. While most of the molecular ions lose either NO˙ + NO˙, or NO˙ + CO, some also lose CO as an initial fragment. One of the fragmentation pathways for 4-methylbenzofurazan-1-oxides involves initial ˙CHO loss. With the annelated benzofurazan-1-oxides (naphtho[1,2-c]furazan oxide and quinolo[3,4-c]furazan oxide), loss of N2O2 is followed by a retro-Diels–Alder elimination of butadiyne or propynenitrile, respectively from the aryne radical cation. In the case of quinolo[5,6-c]furazan oxide, loss of N2O2 from the molecular ion must be followed by substantial rearrangement to enable the observed loss of propynenitrile to take place.  相似文献   

8.
Metal(I) hydrides are eliminated as neutral species in the electron impact ionization mass spectra of copper(II) and palladium(II) complexes of ethylene-N,N′-3-benzoylprop-2-en-2-amine. Deuterium labelling shows that the hydrogen atom of the metal(I) hydride is derived predominantly from the ethylene bridge both for ion source reactions and for metastable ion transitions. Evidence supporting the proposed rationalization for elimination of metal(I) hydride is provided by the observation of an analogous reaction in the mass spectrum of (ethylene-N,N′-salicylaldiminato)copper(II). The mass spectrum of ethylene-d4-N,N′-3-benzoylprop-2-en-2-amine shows an unusual rearrangement to give [C7H5D2]+ ions involving a formal phenyl-to-methylene transfer.  相似文献   

9.
Autoreduction of 2,2,6,6-tetramethyl-1,4-dioxopiperidinium ion to nitroxyl radical in alkaline medium involves a number of parallel and consecutive reactions. The primary products of the reaction of 2,2,6,6-tetramethyl-1,4-dioxopiperidinium with hydroxide ion are three nitroso compounds and N-hydroxy-2,2,6,6-tetramethylpiperidine N-oxide. Isomerization of the nitroso compounds and elimination of acetone from the N-oxide give cyclic hydroxylamines which reduce the initial cation to nitroxyl radical, being oxidized to nitrones.  相似文献   

10.
The mass spectrometric behaviour of 1-aryl-5-(1-acyl-2-dialkylaminovinyl)-1H-tetrazoles was studied, especially using 1-phenyl-5-(1-benzoyl-2-dimethylaminovinyl)-1H-tetrazole 1 and its D-and 15N-labeled derivatives. All tetrazoles investigated showed a clearly observable molecular ion and underwent as the main fragmentation the elimination of nitrogen followed by a number of various subsequent processes. Besides, primary fragments such as [M ? N3?]+ and [M — ArN3]+? were also observed.  相似文献   

11.
Substitution of amino for hydroxyl groups in certain sesquiterpene alcohols has been studied by chemical ionization mass spectrometry using ammonia and ammouia-d3 as the reagent gases, and by mass-analysed ion kinetic energy spectrometry and collision-induced decomposition mass-analysed ion kinetic energy measurements. Depending upon the source conditions and the nature of the substrates, both SNi and SN1 mechanisms have been found to operate. No evidence is obtained for an SN2 mechanism in these compounds. In centdarol and isocentdarol, addition of NH3 to the double bond, followed by elimination of H2O, also contributes to the substitution process. Attack of [NH4]+ on the epoxide function, followed by loss of H2O, appears to lead to the substitution ions in epoxycentdarol, epoxyisocentdarol and epoxyhimachalol.  相似文献   

12.
The use of a versatile N‐heterocyclic carbene (NHC) gold(I) hydroxide precatalyst, [Au(OH)(IPr)], (IPr=N,N′‐bis(2,6‐diisopropylphenyl)imidazol‐2‐ylidene) permits the in situ generation of the [Au(IPr)]+ ion by simple addition of a Brønsted acid. This cationic entity is believed to be the active species in numerous catalytic reactions. 1H NMR studies in several solvent media of the in situ generation of this [Au(IPr)]+ ion also reveal the formation of a dinuclear gold hydroxide intermediate [{Au(IPr)}2(μ‐OH)], which is fully characterized and was tested in gold(I) catalysis.  相似文献   

13.
Abstract

The reaction of sulfimides with hydroxide ion in methanol gave the corresponding sulfoxide (the solvolysis product) and/or the corresponding α-methoxysulfide (the Pummerer type product). The pseudo first order rates for the solvolysis reaction and the Pummerer type reaction were determined using a large excess of potassium hydroxide. The rates of the solvolyses are correlated with [sgrave] values and the values of ρ X = +1.2 and ρ Y = +0.8 were obtained for aryl methyl N-aryl-sulfonylsulfimides (p-XC6H4S(NSO2C6H4Y-p)CH3), and both the activation enthalpy and entropy calculated are δH≠ = 18.8 kcal mol?1 and δS≠ = –23.9 e.u. (PhS(NSO2C6H4CH3-p)CH3), respectively. Hammett correlation with [sgrave] values for the Pummerer type reaction gave ρz = +2.0 for N-aryl-sulfonyltetramethylenesulfimides ((CH2)4SNSO2C6H4Z-p), and the activation enthalpy and entropy are δH≠ = 27.9 kcal mol?1 and δS≠ = +13.3 eu ((CH2)4SNSO2C6H4CH3-p), respectively. All these observations suggest that the solvolysis reaction proceeds via the initial nucleophilic attack of hydroxide ion on the sulfur atom of the sulfimide namely via an S N 2 process at the sulfur atom whereas the Pummerer type reaction proceeds by way of the E 1cb mechanism.  相似文献   

14.

4,5-Dimethyl-1,3-bis(pyridin-2-ylmethyl)-1H-imidazolium chloride (1) was synthesized and characterized by IR and NMR spectroscopy and X-ray diffraction. An attempt to prepare the free tridentate N-heterocyclic carbene pincer ligand by the reaction of 1 with KN(SiMe3)2 resulted in the formation of 1,1′-bis(pyridin-2-ylmethyl)-2,2′-bis(4,5-dimethylimidazole) as a product of dimerization of the target carbene followed by the rearrangement accompanied by the elimination of dipyridylethane.

  相似文献   

15.
Negative ion fast atom bombardment and negative ion chemical ionization tandem mass spectrometry combined with charge inversion reactions are used to confirm that 2-chloro-5nitropyridine reacts with the hydroxide ion by an SN(ANRORC) mechanism in solution to form a ring-opened C5H3N2O 3 ? anion that has the structure of the (M-H)? anion of 2-nitro-4-cyano-2-butenal. The addition of excess hydroxide ion forms the expected 2-oxy5-nitropyridine anion. In the gas phase the same reaction forms only the 2-oxy-5-nitropyridine anion. High energy charge inversion is shown to be an excellent means of differentiating between these isomeric negative ions.  相似文献   

16.
Phenanthrene 9,10-imine ( 1 ) was shown to undergo N-alkylation without aziridine ring cleavage by (a) metallation with sodium methylsulfinylmethide followed by addition of an alkyl halide at −20° (b) reaction of 1 , sodium hydride and the halide in dimethylformamide at 40° (c) treatment of a dichloromethane solution of 1 , the halide and triethylbenzylammonium chloride with aqueous sodium hydroxide under phase transfer conditions. The syntheses of N-isopropyl-, N-butyl-, N-pentyl-, N-allyl- and N-benzylphenanthrene 9,10-imine ( 2–6 ) are described.  相似文献   

17.
Summary The reactions of [Fe(bipym)3]2+ and [Ru(bipym)3]2+ with hydroxide ion in aqueous solution have been followed. The [Ru(bipym)3]2+ species undergoes nucleophilic attack at the ligand to yield [Ru(bipym)2(pyrimidine)(OH)]+ and [HCO2] ion, involving cleavage of one pyrimidyl ring. Intermediates can be observed in the reaction of [Fe(bipym)3]2+ with HO, N3 and SCN. The kinetics of the first reaction have been followed and the results are compared with those known for the reactions of [Fe(bipy)3]2+, [Fe(phen)3]2+ and similar compounds.Part XXIII: P. A. Williams,Transition Met. Chem., 78/84.  相似文献   

18.
5-R-Substituted 1(2)-vinyltetrazoles (R = Ar, Alk, CH2=CH, NH2, H) were synthesized by alkylation of 5-R-tetrazoles with 1,2-dibromoethane in the presence of triethylamine in acetonitrile, followed by elimination of triethylamine hydrobromide. Vinylation of dinuclear substrates, such as bis(1H-tetrazol-5-yl)-methane and 1,3-bis(1H-tetrazol-5-yl)benzene, under analogous conditions gave the corresponding N 1,N 2′- and N 2,N 2′-divinyl derivatives.  相似文献   

19.
The mechanism of elimination of ROH (R = H or CH3) from the ammonium adduct ion, [M+NH4]+, of 1-adamantanol and its methyl ether is examined by using linked-scan metastable ion spectra and by measuring the dependence of the peak intensity ratio [M+NH4]+/[M+NH4? ROH]+ on ammonia pressure. For 1-adamantanol both SNi and SN1 reactions are suggested in metastable ion decomposition, while only the SN1 mechanism is operative in the ion source. For 1-adamantanol methyl ether the SN1 reaction predominates both in metastable ion decomposition and in the ion source reaction.  相似文献   

20.
Degradation via hydrolysis is among the main transformation pathways and particularly for N‐methylcarbamates. Carbamate pesticide hydrolysis is known to proceed through alkaline catalysis, with reaction of the hydroxide ion with the carbonyl function or with abstraction of hydrogen in the α position with respect to the carbonyl. This reaction leads to the formation of methylamine and corresponding phenol. In this respect, the reaction kinetics of 4‐bromo‐3,5‐dimethylphenyl N‐methylcarbamate (BDMC) hydrolysis have been investigated in alkaline solution using a spectrophotometric technique and reversed phase liquid chromatography. The kinetic constants were determined following a proposed pseudo–first‐order kinetic model. The positive activation entropy ΔS = +35.73 J mol−1 K−1 and the absence of general base catalysis indicated an unimolecular elimination conjugate base (E1cB) hydrolytic mechanism involving the formation of methyl isocyanate. This result was confirmed by the fact that BDMC fits well into brönsted and Hammett lines, obtained for a series of substituted N‐methylcarbamate whose decomposition in aqueous media was established to follow an E1cB mechanism.  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