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 34 毫秒
1.
Highest occupied and lowest unoccupied molecular orbital energies, formation energies, bond lengths and FTIR spectra all suggest competing CL-20 degradation mechanisms. This second of two studies investigates recalcitrant, toxic, aromatic CL-20 intermediates that absorb from 370 to 430 nm. Our earlier study (Struct. Chem., 15, 2004) revealed that these intermediates were formed at high OH(-) concentrations via the chemically preferred pathway of breaking the C-C bond between the two cyclopentanes, thereby eliminating nitro groups, forming conjugated pi bonds, and resulting in a pyrazine three-ring aromatic intermediate. In attempting to find and make dominant a more benign CL-20 transformation pathway, this current research validates hydroxylation results from both studies and examines CL-20 transformations via photo-induced free radical reactions. This article discusses CL-20 competing modes of degradation revealed through: computational calculation; UV/VIS and SF spectroscopy following alkaline hydrolysis; and photochemical irradiation to degrade CL-20 and its byproducts at their respective wavelengths of maximum absorption.  相似文献   

2.
Combining computer chemistry calculation with experimental verification is useful both in proving concepts and what is chemically possible. Computational predictions, using MOPAC quantum mechanical and classical force field mechanics, were used to investigate most likely first-tier intermediates of cyclic and cage cyclic nitramines—comparing bond lengths and angles, heats of formation, steric energy, dipole moments, solvent accessibility and electrostatic potential surfaces, partial charges, and Highest Occupied Molecular Orbitals/Lowest Unoccupied Molecular Orbitals (HOMO/LUMO) energies. Two competing modes of degradation are summarized: through addition of hydroxide ions and through addition of photo-induced free radicals. UV/VIS measured concentrations and followed the course of reactions. FTIR followed CL-20 degradation through alkali hydrolysis, where FTIR measurements verified theoretical predictions.  相似文献   

3.
陈丽涛  刘婷  陈加藏 《化学学报》2008,66(10):1187-1195
在B3LYP/LanL2DZ水平上, 计算研究了Co2+, Ni2+, Cu2+, Zn2+, Cd2+与乙二胺四乙酸(EDTA)六配位模式下配合物的结构和性质. 除Cu(EDTA)2-的M—O(5)受Jahn-Teller效应影响明显拉长外, 配位键长M—N(1)和M—O(5)按 Cu2+<Ni2+<Co2+<Zn2+<Cd2+的顺序依次增长, 配位键长M—O(3)按Cu2+<Zn2+<Ni2+<Co2+<Cd2+的次序依次增长. 自然键轨道(NBO)分析表明, 氮、氧的非键电子与金属空轨道的相互作用是配体与金属离子配位的主要作用方式. 通过对N(1)—C(7), N(1)—C(9), N(1)—C(15)键长和NAO键级的分析, 在分子水平上阐明了EDTA在与金属离子配位前后发生首步降解, 其产物存在差异的实验事实. 依据热力学原理并兼顾自洽反应场(SCRF)的IEFPCM模型, 我们设计了金属离子与EDTA在水溶液中的反应途径和热力学循环. 结果表明, 金属离子与EDTA的结合能(即配位稳定性)依次为: Cd2+<Zn2+<Co2+<Ni2+<Cu2+, 金属离子的水合吉布斯自由能计算值与实验值大致吻合, 而且上述目标金属配合物的络合吉布斯自由能的递变规律与实验一致. 基于气相优化结构进行了振动频率计算, 并对部分重要的振动峰作了归属指认. 结果表明, 随着配位稳定性的减弱, M(EDTA)2-具有红外活性的金属敏感性振动峰ν(M—N)和ν(M—O)的峰位依次红移.  相似文献   

4.
The Acinetobacter sp. strain M isolated from a contaminated soil sample in Jiangsu Province of China was found to be able to degrade perfluorooctane sulfonamide (PFOSA) effectively. Fluoride anion (F?) released from PFOSA degradation was detected by ion chromatography, and showed positive correlation to the growth curve of Acinetobacter sp. strain M. The PFOSA degradation efficiency of strain M was approximately 27 %, as assessed by GC analysis. It was shown that enzymes localized outside of cells of Acinetobacter sp. strain M catalyzed the degradation of PFOSA. This further indicates a possibly new (multi‐step/pathway) mechanism for PFOSA degradation. It revealed that the extracellular enzyme of the Acinetobacter strain M preferentially cleaves carbon‐carbon and carbon‐fluorine bonds instead of destroying the carbon‐sulfur bond. The growth condition for Acinetobacter sp. strain M was optimized at 30 °C and pH 7.0 in the presence of 2000 mg L?1 of PFOSA and 0.5 % (v/v) of Tween‐20. The optimal PFOSA degradation time was found to be 12 h, with a degradation efficiency of 76 % by extracellular enzymes in strain M as determined by GC analysis. The result may provide potential applications for biodegradition of perfluoro organic compounds, such as derivatives of perfluorooctane (C8).  相似文献   

5.
Oxidation of Li/X phosphinidenoid complex 2 , obtained via selective deprotonation from the P‐H precursor 1 , with [Ph3C]BF4 led to the formation of two P‐F substituted diorganophosphane complexes 6 , 7 ; the latter tautomer 7 formed via H‐shift from 6 . In contrast, oxidation of 2 with [(p‐Tol)3C]BF4 led to three major and one minor intermediates at low temperature, which we tentatively assign to two pairs of P‐C atropisomers 10a , a′ and 10c , c′ and which differ by the relative orientations of their CH(SiMe3)2 and W(CO)5 groups. Conversion of all isomers led finally to complex 11 having a ligand with a long P? C bond to the central trityl* carbon atom, firmly established by single‐crystal X‐ray analysis. DFT calculations at the B3LYP/def2‐TZVPP//BP86/def2‐TZVP level of theory on real molecular entities revealed the structures of the in situ formed combined singlet diradicals ( 4 + 5 and 5 + 9 ) and the nature of intermediates on the way to the final product, complex 11 . Remarkable is that all isomers of 11 possess relative energies in the narrow energy regime of about 20 kcal mol?1. A preliminary study revealed that complex 11 undergoes selective P? C bond cleavage at 75 °C in toluene solution.  相似文献   

6.
Undirected C(sp3)?H functionalization reactions often follow site‐selectivity patterns that mirror the corresponding C?H bond dissociation energies (BDEs). This often results in the functionalization of weaker tertiary C?H bonds in the presence of stronger secondary and primary bonds. An important, contemporary challenge is the development of catalyst systems capable of selectively functionalizing stronger primary and secondary C?H bonds over tertiary and benzylic C?H sites. Herein, we report a Cu catalyst that exhibits a high degree of primary and secondary over tertiary C?H bond selectivity in the amidation of linear and cyclic hydrocarbons with aroyl azides ArC(O)N3. Mechanistic and DFT studies indicate that C?H amidation involves H‐atom abstraction from R‐H substrates by nitrene intermediates [Cu](κ2N,O‐NC(O)Ar) to provide carbon‐based radicals R. and copper(II)amide intermediates [CuII]‐NHC(O)Ar that subsequently capture radicals R. to form products R‐NHC(O)Ar. These studies reveal important catalyst features required to achieve primary and secondary C?H amidation selectivity in the absence of directing groups.  相似文献   

7.
Methylation and demethylation of mercury compounds are two important competing processes that control the net production of highly toxic mercury alkyls, methylmercury (MeHg+) and dimethylmercury (Me2Hg), in environment. Although the microbial and the photochemical methylation and demethylation processes are well studied in recent years but the chemical methylation and demethylation processes have not been studied well. Herein, we report for the first time that the CuSe nanosheet has remarkable ability to activate the highly inert Hg?C bonds of various MeHg+ and Me2Hg compounds at room temperature (21 °C). It facilitates the conversion of MeHg+ into Me2Hg in the absence of any proton donors. Whereas, in the presence of any proton source, it has unique ability to degrade MeHg+ into CH4 and inorganic mercury (Hg2+). Detailed studies revealed that the relatively fast Hg?C bond cleavage was observed in case of MeHgSPh or MeHgI in comparison to MeHgCl, indicating that the Hg?C bond in MeHgCl is relatively inert in nature. On the other hand, the Hg?C bond in Me2Hg is considered to be exceedingly inert and, thus, difficult to cleave at room temperature. However, CuSe nanosheets showed unique ability to degrade Me2Hg into CH4 and Hg2+, via the formation of MeHg+, under acidic conditions at room temperature. DFT calculations revealed that the Hg?C bond activation occurs through adsorption on the surface of (100)‐faceted CuSe nanosheets.  相似文献   

8.
13,14-Dihydroxy-8,11,13-podocarpatrien-7-one (1) and a series of ring C aromatic diterpene derivatives were synthesised from (+)-manool (4) and evaluated for their cytotoxic, leishmanicidal and trypanocidal activities. Our results indicated that compound 1 and other podocarpane-type intermediates are cytotoxic. Cleavage of C6–C7 bond of compound 7 improved cytotoxic activity, indicating that, in particular, the 6,7-seco-podocarpane-type compound 20 might serve as a lead compound for further development.  相似文献   

9.
Degradation of ethyl tert‐butyl ether (ETBE) with UV/TiO2 was studied by solid‐phase microextraction and gas chromatography‐mass spectrometry. The complete removal of 0.1 g L?1 of ETBE was achieved after 20 h of treatment. Factors such as pH of the system, catalyst and substrate concentration, and the presence of anions influenced the degradation rate. Establishment of the degradation pathway was made possible by a thorough analysis of the reaction mixture, which identified the main intermediate products generated. The possible degradation pathways were proposed and discussed in this research. The attack on the C–H bond in ETBE by ·OH forms an alkyl radical, which consequently produces a peroxyl radical upon reaction with oxygen. Peroxyl radicals react with one another and produce an alkoxy radical. The β‐bond fragmentation of the alkoxy radical produces different intermediates.  相似文献   

10.
The 2-exo-substituted fenchol derivatives 1 – 7 , easily prepared from (−)-fenchone in good-to-excellent yields, were pyrolyzed by dynamic gas-phase thermo-isomerization (DGPTI). At temperatures of ca. 620°, the substrates with a hydroxyallyl ( 1 – 4 ) or a hydroxypropargyl moiety ( 6 ) underwent an initial retro-ene reaction under cleavage of the C(2) C(3) bond to form enol-ene intermediates with no loss of optical activity. These intermediates then experience either tautomerization to the corresponding α,β-unsaturated ketones or subsequent Conia rearrangement under one-carbon ring expansion of the fenchone system to a bicyclo[3.2.1]octane framework. In the case of the isopropenyl substrate 3 , the sterically crowded Conia product underwent a new type of ‘deethanation’ reaction by stepwise loss of two Me radicals, giving rise to the thermodynamically favored enone 21 . A similar relaxation behavior was observed in the case of the ethynyl substrate 6 , which showed a remarkable 1,3-Me shift after the Conia reaction, leading to the α,β-unsaturated cyclic ketone 25 . The homolytic cleavage of the weakest single bond in 1 – 3 turned out to be a competing reaction pathway. Intramolecular H-abstraction within the generated diradical intermediates produced the monocyclic ketones 8, 16 , and 19 , besides the products obtained by tautomerization and Conia reaction. In contrast, a Ph substituent at C(2) in 7 allowed only the passage through a diradical species to provide phenone 26 , which was converted by regioselective Baeyer–Villiger oxidation to the optically active cyclopentanol 29 . Both reaction channels, the domino retro-ene–Conia rearrangement and the diradical-promoted H-transfer, have been shown to proceed highly stereoselectively. The absolute configuration of the newly formed stereogenic centers in all compounds was assigned by 1H-NOE experiments. The reaction mechanism of the novel domino retro-ene–Conia reaction was established by both a series of 2H- and 13C-labeling experiments, as well as by a detailed computational analysis.  相似文献   

11.
Catalytic water oxidation at Ir (OH)+ ( Ir =IrCp*(Me2NHC), where Cp*=pentamethylcyclopentadienyl and Me2NHC=N,N′‐dimethylimidazolin‐2‐ylidene) can occur through various competing channels. A potential‐energy surface showing these various multichannel reaction pathways provides a picture of how their importance can be influenced by changes in the oxidant potential. In the most favourable calculated mechanism, water oxidation occurs via a pathway that includes four sequential oxidation steps, prior to formation of the O?O bond. The first three oxidation steps are exothermic upon treatment with cerium ammonium nitrate and lead to formation of Ir V(?O)(O . )+, which is calculated to be the most stabile species under these conditions, whereas the fourth oxidation step is the potential‐energy‐determining step. O?O bond formation takes place by coupling of the two oxo ligands along a direct pathway in the rate‐limiting step. Dissociation of dioxygen occurs in two sequential steps, regenerating the starting material Ir (OH)+. The calculated mechanism fits well with the experimentally observed rate law: v=kobs[ Ir ][oxidant]. The calculated effective barrier of 24.6 kcal mol?1 fits well with the observed turnover frequency of 0.88 s?1. Under strongly oxidative conditions, O?O bond formation after four sequential oxidation steps is the preferred pathway, whereas under milder conditions O?O bond formation after three sequential oxidation steps becomes competitive.  相似文献   

12.
The behaviour of allylbutyltin dichloride in water, water–ethanol and water–hexane media, under either homogeneous or heterogeneous conditions, has been studied. 1,3-Diallyl-1,3-dibutyl-1,3-dichlorodistannoxane, butyltin di(hydroxy)chloride and butyltin trichloride arise from the solvolytic, acid–base and degradation processes. The degradation process involving the cleavage of the tin–carbon allyl bond has been interpreted to occur via an intramolecular reaction at the expense of the cation [Bu(CH2=CHCH2)Sn(OH)(H2O)n]+. The mechanistic pathway is ascribable to an internal interaction of the electrophilic cation with a bonded water molecule. This mechanistic proposal may be of some help with understanding of the chemical degradation of diorganotin derivatives in aquatic environments.  相似文献   

13.
For the first time, a highly regioselective intermolecular sulfonylamidation unactivated secondary Csp3?H bond has been achieved using IrIII catalysts. The introduced N,N’‐bichelating ligand plays a crucial role in enabling iridium–nitrene insertion into a secondary Csp3?H bond via an outer‐sphere pathway. Mechanistic studies and density functional theory (DFT) calculations demonstrated that a two‐electron concerted nitrene insertion was involved in this Csp3?H amidation process. This method tolerates a broad range of linear and branched‐chain N‐alkylamides, and provides efficient access to diverse γ‐sulfonamido‐substituted aliphatic amines.  相似文献   

14.
Mechanism of substrate oxidations with hydrogen peroxide in the presence of a highly reactive, biomimetic, iron aminopyridine complex, [FeII(bpmen)(CH3CN)2][ClO4]2 ( 1 ; bpmen=N,N'‐dimethyl‐N,N'‐bis(2‐pyridylmethyl)ethane‐1,2‐diamine), is elucidated. Complex 1 has been shown to be an excellent catalyst for epoxidation and functional‐group‐directed aromatic hydroxylation using H2O2, although its mechanism of action remains largely unknown. 1 , 2 Efficient intermolecular hydroxylation of unfunctionalized benzene and substituted benzenes with H2O2 in the presence of 1 is found in the present work. Detailed mechanistic studies of the formation of iron(III)–phenolate products are reported. We have identified, generated in high yield, and experimentally characterized the key FeIII(OOH) intermediate (λmax=560 nm, rhombic EPR signal with g=2.21, 2.14, 1.96) formed by 1 and H2O2. Stopped‐flow kinetic studies showed that FeIII(OOH) does not directly hydroxylate the aromatic rings, but undergoes rate‐limiting self‐decomposition producing transient reactive oxidant. The formation of the reactive species is facilitated by acid‐assisted cleavage of the O? O bond in the iron–hydroperoxide intermediate. Acid‐assisted benzene hydroxylation with 1 and a mechanistic probe, 2‐Methyl‐1‐phenyl‐2‐propyl hydroperoxide (MPPH), correlates with O? O bond heterolysis. Independently generated FeIV?O species, which may originate from O? O bond homolysis in FeIII(OOH), proved to be inactive toward aromatic substrates. The reactive oxidant derived from 1 exchanges its oxygen atom with water and electrophilically attacks the aromatic ring (giving rise to an inverse H/D kinetic isotope effect of 0.8). These results have revealed a detailed experimental mechanistic picture of the oxidation reactions catalyzed by 1 , based on direct characterization of the intermediates and products, and kinetic analysis of the individual reaction steps. Our detailed understanding of the mechanism of this reaction revealed both similarities and differences between synthetic and enzymatic aromatic hydroxylation reactions.  相似文献   

15.
A series of azo-ester linked mesogen containing liquid crystalline acrylate compounds C1-C6 having different terminal groups (–F, –Cl, –Br, –OCH3, –OC2H5 and –OC3H7) were successfully synthesised and characterised. The chemical structure, purity, thermal stability, mesophase behaviour and optical property of the synthesised compounds were investigated by different instrumental techniques. X-ray crystal structure showed that compounds C1, C4 and C5 exhibited more stable E configuration with two bulky group in the opposite side of the N=N double bond motifs. The fluoro-substituted derivative (C1) is connected by the R12(5) type of C–H…O hydrogen bond motifs whereas the molecules of C4, and C5 are connected to each other by means cyclic R22(8) type of C–H…O hydrogen bond motifs. Thermogravimetric study revealed that the investigated compounds exhibited excellent thermal stability. All the compounds showed enantiotropic liquid crystal (LC) phase behaviour and the mesophase formation was greatly influenced by the terminal substituents. Alkoxy (–OCH3, –OC2H5 and –OC3H7) substituted compounds exhibited greater mesophase stability than those of halogen (–F, –Cl and –Br) terminated derivatives. UV-vis spectroscopic study revealed that the investigated compounds exhibited a broad absorption band around 300–420 nm with absorption maximum (λmax) of nearly 370 nm.  相似文献   

16.
A reaction pathway for the rearrangement-dehydrogenation of cis-1-cyclohexyl-2-phenyl-3-benzoylaziridine into 2-cyclohexylamino-3-phenylindenone can now be suggested. Furthermore, a competing degradation pathway involving C? C bond scission accounts for the major product in these reactions and leads to ω-cyclohexylaminoacetophenone and benzaldehyde. Observed also is the fact that trans-1-cyclohexyl-2-phenyl-3-benzoylaziridine fails to undergo the rearrangement-dehydrogenation reaction.  相似文献   

17.
The water-assisted hydrolytic deamination mechanism of adenine was studied using density functional method at B3LYP/6-311G(d,p) level. Intrinsic reaction coordinate (IRC) calculations were performed on the transition states to verify whether it is the real transition states that connect the corresponding intermediates. Single-point calculations were carried out on the previous optimized geometries obtained during IRC calculations. The activation energies have also been calculated using G3MP2//B3LYP/6-311G(d,p) method. The water molecules attack the adenine and a tetrahedral intermediate forms. Then, two different intermediates have been obtained through different bond rotations. In pathway a, the second water molecule takes part in the formation of transition state and acts as a bridge to transfer hydrogen atom, while in pathway b, the second water molecule does not involve in the creation of transition state and only acts as a medium. The energy barriers are 23.40 and 37.17 kcal/mol for pathways a and b, respectively. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

18.
Using chloroacetic acid, p‐hydroxyl aromatic aldehydes 1 and aromatic diamines 3 as starting materials, novel bisbenzimidazoles 5 with unsymmetric structure were synthesized via aryloxyacetic acid intermediates 2 . Four new intermediates 4 and ten target molecules 5 were characterized by FTIR, 1H NMR, 13C NMR, MS and elemental analysis. Different synthetic methods, including one‐pot synthesis and intermittent microwave promotion, were investigated. The research provides a new method and idea for the synthesis of bisbenzimidazoles.  相似文献   

19.
Abstract

A simple and eco-compatible synthesis of podands as 4a-i is performed using amines (3a-c ), CS2, and dichlordiamides (DCDs) (2a, c ) in the absence of a catalyst in water. Three reacting DCDs (2a-c ) were obtained in the reaction of aromatic diamines (1a, c ) with chloroacetyl chloride. Dynamic NMR spectroscopic data of three series of podands (4a-c , 4d-f , and 4g-i ) are discussed and their free energies of activations are calculated (ΔG c s) at coalescence temperatures. The ΔG c of these podands were attributed to conformational isomerization in the range of 14.7–17 kcal/mol due to rotation and resonance effects about the thioamide C?N bond.

[Supplementary materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfer, and Silicon and the Related Elements for the following free supplemental files: Additional figures.]  相似文献   

20.
Gemifloxacin mesylate (GFM), chemically (R,S)‐7‐[(4Z)‐3‐(aminomethyl)‐4‐(methoxyimino)‐1‐pyrrolidinyl]‐1‐cyclopropyl‐6‐fluoro‐1,4‐dihydro‐4‐oxo‐1,8‐naphthyridine‐3‐carboxylic acid methanesulfonate, is a synthetic broad‐spectrum antibacterial agent. Although many papers have been published in the literature describing the stability of fluorquinolones, little is known about the degradation products of GFM. Forced degradation studies of GFM were performed using radiation (UV‐A), acid (1 mol L?1 HCl) and alkaline conditions (0.2 mol L?1 NaOH). The main degradation product, formed under alkaline conditions, was isolated using semi‐preparative LC and structurally elucidated by nuclear magnetic resonance (proton – 1H; carbon – 13C; correlate spectroscopy – COSY; heteronuclear single quantum coherence – HSQC; heteronuclear multiple‐bond correlation – HMBC; spectroscopy – infrared, atomic emission and mass spectrometry techniques). The degradation product isolated was characterized as sodium 7‐amino‐1‐pyrrolidinyl‐1‐cyclopropyl‐6‐fluoro‐1,4‐dihydro‐4‐oxo‐1,8‐naphthyridine‐3‐carboxylate, which was formed by loss of the 3‐(aminomethyl)‐4‐(methoxyimino)‐1‐pyrrolidinyl ring and formation of the sodium carboxylate. The structural characterization of the degradation product was very important to understand the degradation mechanism of the GFM under alkaline conditions. In addition, the results highlight the importance of appropriate protection against hydrolysis and UV radiation during the drug‐development process, storage, handling and quality control. Copyright © 2015 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