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 41 毫秒
1.
The reactions of oximes to amides, known as the Beckmann rearrangement, may undergo fragmentation to form carbocations + nitriles when the migrating groups have reasonable stability as cations. The reactions of oxime sulfonates of 1-substituted-phenyl-2-propanone derivatives (7-X) and related substrates (8-X, 9a-X) in aqueous CH(3)CN gave both rearrangement products (amides) and fragmentation products (alcohols), the ratio of which depends on the system; the reactions of 7-X gave amides predominantly, whereas 9a-X yielded alcohols as the major product. The logk-logk plots between the systems gave excellent linear correlations with slopes of near unity. The results support the occurrence of path bifurcation after the rate-determining TS of the Beckmann rearrangement/fragmentation reaction, which has previously been proposed on the basis of molecular dynamics simulations. It was concluded that path-bifurcation phenomenon could be more common than thought and that a reactivity-selectivity argument based on the traditional TS theory may not always be applicable even to a well-known textbook organic reaction.  相似文献   

2.
[reaction: see text] The BF(3)-catalyzed rearrangement of cyclobutene-fused m- and p-substituted diarylhomobenzoquinones exclusively gave the keto-alcohols via a Wagner-Meerwein vinyl-anion migration followed by the annulation of a delta-located endo-aryl group. The Hammett treatments for the endo/exo substituent effects, as well as the kinetic solvent effects, indicated that this reaction proceeds through the concerted S(N)2-like mechanism involving a rate-determining endo-aryl-assisted transition state.  相似文献   

3.
The mechanisms for the major fragmentations obtained with selected substituted uracils are discussed. Interpretation of data was facilitated by use of metastable peaks, high-resolution data, and low-voltage spectra. The major fragmentation obtained with N-alkyl substituted uracils, when the alkyl group contains 2 or more carbons, is due to cleavage of the alkyl substituent. This cleavage is accompanied by a rearrangement of 1 or 2 hydrogens from the alkyl group to the uracil ring. Possible mechanisms for the rearrangements are discussed. It was found that the molecular ion of 1- and 3-alkyl substituted uracils (where the alkyl group has 2 or more carbons) does not undergo the expected ‘retro Diels-Alder Reaction’. Instead, the odd-electron ion formed by loss of the alkyl substituent with a single hydrogen rearrangement undergoes this reaction (loses HNCO). Since it is formed as a secondary reaction product, the relative abundance of the ‘retro Diels-Alder’ fragment is low compared to what is obtained in the spectra of the simple uracils. The ‘retro Diels-Alder Reaction’ can be used to differentiate between 2- and 4-thiouracils, and between 1- and 3-methyl and phenyl substituted uracils. It was found that 1- and 3-alkyl substituted uracils (alkyl group of 2 or more carbons) can be differentiated by the mass of the M-alkyl fragment since the 3-substituted compounds give predominantly a double hydrogen rearrangement and the 1-substituted compound gives mainly a single hydrogen rearrangement. In addition the intensity of the molecular ion, relative to the M-alkyl ion, is considerably stronger in the 1-substituted uracils.  相似文献   

4.
The principal pathways of the fragmentation of (3-quinuclidinyl)diaryl(heteryl)carbinols that involve cleavage of the quinudidine-carbinol C-C bond and the bridge bond in the quinuclidine ring containing the substituent were studied. In addition to the indicated fragmentation pathways, fragmentation proceeding with opening of the bridge bond of quinuclidine that does not contain a substituent is observed. The rearrangement of the molecular ion that precedes fragmentation via the indicated pathway is examined.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 8, pp. 1078–1082, August, 1990.The authors thank V. G. Zhil'nikov (A. N. Severtsov Institute of the Evolutionary Morphology and Ecology of Animals) for recording the high-resolution mass spectra.  相似文献   

5.
In the presence of a catalytic amount of 3-butyl-4-methylthiazolium bromide, the reaction of benzaldehydes with azobenzene in dichloromethane containing octanethiol and Et(3)N gave the corresponding S-octyl thiobenzoates in good yields. The thiolesters were produced by trapping of the 2-benzoylthiazolium salts with the thiol, which were generated through the azobenzene oxidation of the active aldehydes. This is the first example for the thiolester formation mimicking the function of the pyruvate dehydrogenase complex. An electron-withdrawing substituent at the 4-position of benzaldehyde enhanced the reaction rate. The effect of benzaldehyde substituents on the reaction rate was examined quantitatively on the basis of kinetic measurements, leading to a nonlinear correlation of log(k(obs)) with Hammett's substituent constants (sigma). The origin of the nonlinear Hammett plot was interpreted in terms of a shift in the rate-determining step of the multistep reaction with change of the electronic nature of substituent. Further support for this assumption was given by the observation that the reaction constant (rho) of the Hammett plot for the azobenzene substituent effect on the oxidation rate of 4-bromobenzaldehyde was much smaller than that of 4-cyanobenzaldehyde.  相似文献   

6.
本文研究了新近合成的十三个2-芳基-5(2-苯基-4-喹啉基)-1,3,4-噁二唑衍生物的电子轰击(EI-70eV)质谱。应用亚稳离子和质量准确测定方法研究了这些化合物在EI-70eV条件下的断裂途径。归纳出以环状结构的分解和重排为特征的几种主要断裂方式。并用Hammett常数σ拟合了取代基效应与裂解反应的关系。  相似文献   

7.
The Schmidt reaction is the acid-catalyzed analogue of the Curtius reaction and is extensively used in organic synthesis. In this work, the mechanism of this reaction has been explored using DFT calculations at the B3LYP/6-311+G(d,p) level. Protonated formyl azide may undergo rearrangement to the product, protonated isocyanic acid, with simultaneous extrusion of molecular nitrogen (concerted mechanism), or undergo rearrangement to the anti conformer, followed by removal of nitrogen to form the nitrenium ion, which then rearranges to the final product, protonated isocyanic acid (step-wise mechanism). Like the Curtius reaction, it is found that the concerted pathway is definitely preferred. The key role of acidification in decreasing the overall energy barrier is more highlighted in case of phenyl substitution, with negligible effect on the lower homologues. For methoxy and amine substituents, there is very little difference in the activation energies of the concerted and step-wise reactions, with the former being still slightly preferred. Unlike the parent compound, the rearrangement of substituted nitrenium ion in some cases involves side reactions like C-H insertion and cyclization.  相似文献   

8.
The combination of thiamin and benzaldehyde can produce benzoin but also destroys thiamin. The destruction comes from fragmentation of the conjugate of thiamin and benzaldehyde undergoing a process that produces a phenyl thiazole ketone and pyrimidine. The key step in this process is cleavage of the C-N bond between the heterocycles, which occurs by an unknown mechanism. Enzymes that utilize similar intermediates do not fragment the cofactor although fragmentation is inherent to the structure. To analyze the nature of the C-N cleavage step, the rates of fragmentation of a series of phenyl-substituted N1'-methyl-2-(1-hydroxybenzyl)thiamin derivatives were determined under two sets of conditions: (1) where proton transfer in the step prior to C-N bond breaking is rate-determining and (2) where C-N bond breaking is rate-determining. The resulting rho values are 1.6 and 1.8, respectively, leading to the conclusion that C-N cleavage is insensitive to substituent effects. On the basis of these results, we conclude that cleavage occurs by a facile process that resembles the outcome of a [1,5] sigmatropic rearrangement. An enzyme may avoid the fragmentation by holding the intermediate in a conformation that prevents such a process, allowing the normal catalytic process to proceed.  相似文献   

9.
Tin tetrachloride-catalyzed glycosidation of persilylated nucleobases with acetate donor 6 in CH(2)Cl(2) solution followed by deprotection gave rise very predominantly to alpha-spironucleosides. These stereochemical assignments stem from the determination of NOE interactions and an X-ray crystallographic analysis of the latter product. Computational studies revealed that these results are consistent with the fact that the C5' substituent shields the beta-face of the oxonium ion involved in the coupling reaction while the C3' substituent is projected away from the alpha-underside. Attack from the more open direction is therefore kinetically favored. Entirely comparable calculations suggested that donor 19 should behave comparably. Experimentation involving this donor gave results consistent with this model although more equitable alpha/beta spironucleoside product ratios were seen when acetonitrile was employed as the reaction medium.  相似文献   

10.
The correlation of anion structure with the fragmentation behavior of deprotonated nitrobenzenesulfonylamino acids was investigated using tandem mass spectrometry, isotopic labeling and computational methods. Four distinct fragmentation pathways resulting from the collision‐induced dissociation (CID) of deprotonated 2‐nitrobenzenesulfonylglycine (NsGly) were characterized. The unusual loss of the aryl nitro substituent as HONO was the lowest energy process. Subsequent successive losses of CO, HCN and SO2 indicated that an ortho cyclization reaction had accompanied loss of HONO. Other pathways involving rearrangement of the ionized sulfonamide group, dual bond cleavage and intramolecular nucleophilic displacement were proposed to account for the formation of phenoxide, arylsulfinate and arylsulfonamide product ions at higher collision energies. The four distinct fragmentation pathways were consistent with precursor–product relationships established by CID experiments, isotopic labeling results and the formation of analogous product ions from 2,4‐dinitrobenzenesulfonylglycine and the Ns derivatives of alanine and 2‐aminoisobutyric acid. The computations confirmed a low barrier for ortho cyclization with loss of HONO and feasible energetics for each reaction step in the four pathways. Computations also indicated that three of the fragmentation pathways started from NsGly ionized at the carboxyl group. Overall, the pathways identified for the fragmentation of the NsGly anion differed from processes reported for anions containing a single functional group, demonstrating the importance of functional group interactions in the fragmentation pathways of multifunctional anions. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

11.
The gas-phase dissociation of phosphorylated peptides was modelled using a combination of quantum mechanics and the Rice-Ramsperger-Kassel-Marcus theory. Potential energy surfaces and unimolecular reaction rates for several low-energy fragmentation and rearrangement pathways were estimated, and a general mechanism was proposed. The neutral loss of the phosphoric acid was mainly an outcome of the intramolecular nucleophilic substitution mechanism. The mechanism involves a nucleophilic attack of the phosphorylated amino acid N-terminal carbonyl oxygen on β-carbon, yielding a cyclic five-membered oxazoline product ion. Regardless of the proton mobility, the pathway was charge directed either by a mobile proton or by a positively charged side chain of some basic residue. Although the mechanistic aspects of the phosphate loss are not influenced by the proton mobility environment, it does affect ion abundances. Results suggest that under the mobile proton environment, the interplay between phosphoric acid neutral loss product ion and backbone cleavage fragments should occur. On the other hand, when proton mobility is limited, neutral loss product ion may predominate. The fragmentation dynamics of phosphoserine versus phosphothreonine containing peptides suggests that H(3)PO(4) neutral loss from phosphothreonine containing peptides is less abundant than that from their phosphoserine containing analogs. During the low-energy CID of phosphorylated peptides in the millisecond time range, typical for ion trap instruments, a phosphate group rearrangement may happen, resulting in an interchange between the phosphorylated and the hydroxylated residues. Unimolecular dissociation rate constants imply the low abundance of such scrambled product ions.  相似文献   

12.
Rate constants were measured for the oxidative chlorodehydrogenation of (R,S)-2-phenoxypropanoic acid and nine ortho-, ten para- and five meta-substituted derivatives using (R,S)-1-chloro-3-methyl-2,6-diphenylpiperidin-4-one (NCP) as chlorinating agent. The kinetics was run in 50% (v/v) aqueous acetic acid acidified with perchloric acid under pseudo-first-order conditions with respect to NCP at temperature intervals of 5 K between 298 and 318 K, except at the highest temperature for the meta derivatives. The dependence of rate constants on temperature was analyzed in terms of the isokinetic relationship (IKR). For the 20 reactions studied at five different temperatures, the isokinetic temperature was estimated to be 382 K, which suggests the preferential involvement of water molecules in the rate-determining step. The dependence of rate constants on meta and para substitution was analyzed using the tetralinear extension of the Hammett equation. The parameter lambda for the para/meta ratio of polar substituent effects was estimated to be 0.926, and its electrostatic modeling suggests the formation of an activated complex bearing an electric charge near the oxygen atom belonging to the phenoxy group. A new approach is introduced for examining the effect of ortho substituents on reaction rates. Using IKR-determined values of activation enthalpies for a set of nine pairs of substrates with a given substituent, a linear correlation is found between activation enthalpies of ortho and para derivatives. The correlation is interpreted in terms of the selectivity of the reactant toward para- or ortho-monosubstituted substrates, the slope of which being related to the ortho effect. This slope is thought to be approximated by the ratio of polar substituent effects from ortho and para positions in benzene derivatives. Using the electrostatic theory of through-space interactions and a dipole length of 0.153 nm, this ratio was calculated at various positions of a charged reaction center along the benzene C1-C4 axis, being about 2.5 near the ring and decreasing steeply with increasing distance until reaching a minimum value of -0.565 at 1.3 nm beyond the aromatic ring. Activation enthalpies and entropies were estimated for substrates bearing the isoselective substituent in either ortho and para positions, being demonstrated that they are much different from the values for the parent substrate. The electrophilic attack on the phenolic oxygen atom by the protonated chlorinating agent is proposed as the rate-determining step, this step being followed by the fast rearrangement of the intermediate thus formed, leading to products containing chlorine in the aromatic ring.  相似文献   

13.
The kinetic profile of the Newman-Kwart rearrangement has been evaluated using microwave heating. After first demonstrating equivalence between conventional convective heating and microwave heating, data was gathered and analyzed to determine the effects of substituent, solvent, and concentration on the reaction order. Reaction rate constants, Arrhenius constants, and activation energies have been determined. The reaction rate shows strong sensitivity to the substituent and modest sensitivity to the solvent. At high concentrations, the reaction order increases from the previously reported first-order to a mixed first/second-order reaction. Overall, this re-evaluation of the Newman-Kwart rearrangement has shown the reaction rate order to be more complex than previously thought. In addition, microwave heating has proven ideal for the rapid collection of data to facilitate this type of kinetic study.  相似文献   

14.
Moore IF  Kluger R 《Organic letters》2000,2(14):2035-2036
[reaction: see text] The rapid fragmentation of 2-(1-hydroxybenzyl)thiamin (1) is initiated by transfer of a proton from C2alpha to give an enamine. The subsequent irreversible process can be written as a concerted (or stepwise) rearrangement involving migration of the hydroxyl hydrogen to the methylene bridge. An attractive alternative is internal addition of C2alpha to the pyrimidine, generating a carbocation. However, addition of azide to the reaction solution, which could trap the carbocation, has no effect on the rate or products of reaction.  相似文献   

15.
Reactions of diphenyl sulfide with the two oxidants peroxydisulfate and peroxydiphosphate are described. With each oxidant, the reaction is first order in the oxidant and zero order in the substrate. The rate of the reaction is also independent of the effect of the substituent. Hydrogen ion catalyzes the reactions. The lack of inhibition of the reaction rate by added acrylamide rules out the possibility of a radical reaction initiated by the homolysis of the peroxyanions. From the effect of [H+] on the oxidation rates, the active species involved in the reactions have been determined. The redox reaction is essentially proceeding via hydrolysis of the protonated peroxy anions in a rate-determining step, followed by a fast step involving the oxidation of diphenyl sulfide to diphenyl sulfoxide. This has been confirmed by estimating the hydrolyzed products in each case and also by the product analyses.  相似文献   

16.
A previously unreported series of N-(substituted benzalamino)phthalimides was investigated by using the combined techniques of high resolution electron ionization mass spectrometry, metastable decomposition, and collisional activation mass spectrometry. The predominate fragmentation pathway is a McLafferty-type rearrangement. There also occurs, to a lesser extent, a transfer of hydrogen that originates from a substituent remote from the phthalimide moiety and terminates on the phthalimide, The process is interpreted as proceeding via an ion-neutral complex. The effects of substituents on both of the aforementioned fragmentation pathways provide a striking example that gives quantitative evidence for Stevenson’s rule. The substituent effects are responsible for a trend in ion abundance that shows a sharp reversal at approximately the ionization energy of the iminium isomer of the phthalimide molecular ion.  相似文献   

17.
The mass spectral fragmentation modes of various 2-chloro-3,6-diaryl-3,4-dihydro-1,3,2-oxazaphosphorin-2-oxides—a novel ring system—reveal cycloreversion by two pathways. Retro Diels-Alder reaction by a stepwise mechanism is prominent in this system. The relative abundances of the enone and dienophile ions depend on the nature of the substituent attached to the double bond in this ring. Another retro Diels-Alder fragmentation process, involving the loss of PO2Cl from the molecular ion, is preceded by a 1,3-allylic rearrangement and is the major fragmentation mode in the metastable time scale. Further fragmentation of the [M? PO2Cl]+ imine ion seems to occur from cyclic dihydroquinoline intermediates by substitution elimination steps.  相似文献   

18.
The fragmentation behaviour of ten 3,4-disubstituted 1,2,4-oxadiazole-5(4H)-thiones and seven 3,4-disubstituted 1,2,4-thiadiazole-5(4H)-ones studied here confirmed the earlier observations about the partial rearrangement of the former after ionization into the latter before further fragmentation. In the case of eight 3-(substituted phenyl)-4-(p-tolyl)-1,2,4-oxadiazole-5(4H)-thiones the fragmentations reflecting the above-mentioned molecular ion rearrangement show a clear correlation on the substituent sigma values. The electron-withdrawing substituents destabilize the molecular ion, so higher amounts of the rearranged ion [R(1)NCO](+.) are obtained. A good correlation of log[R(1)NCO](+) against sigma was obtained (r = 0.96). Only a satisfactory correlation prevailed for log([R(1)NCO](+)*/[R(1)NCS](+)*) against sigma(r = 0.87).  相似文献   

19.
Direct derivatization of the acyl analogue of platelet-activating factor (acyl-PAF) with heptafluorobutyric anhydride results in replacement of the phosphocholine moiety with a heptafluorobutyryl (HFB) group. Electron capture (EC) mass spectrometric analysis of this compound that makes use of negative ion detection along with subsequent accurate mass measurement and tandem mass spectrometry studies revealed that in addition to expected fragmentation due to losses of elements of HF, ketene, and/or acetic acid, there is a rearrangement reaction between the HFB group and the substituent on carbon-2 of the glycerol backbone. For 2-acetyl isomers, this fragmentation yields a characteristic ion at m/z 237; for 1-acetyl isomers, the analogous ion is observed at [M-135], along with a corresponding carboxylate anion. The use of the HFB derivative is invaluable for analysis of PAF homologues and analogues because it provides detailed structural information in combination with the high sensitivity of a gas chromatography combined with EC-mass spectrometry assay.  相似文献   

20.
A series of alkylphenylketones has been examined. Whenever the alkyl chain is three or more carbon atoms long, the well-known McLafferty rearrangement occurs with elimination of the elements of a neutral olefin. The further fragmentation of the ion formed in this rearrangement reaction has been studied using the technique of ion kinetic energy (IKE) spectroscopy. The measured release of kinetic energy has been used to show that the rearrangement ion has an enolic type structure.  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