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 15 毫秒
1.
The acid‐catalyzed hydrolysis of a series of 5‐substituted‐1H,3H‐2,1,3‐benzothiadiazole 2,2‐dioxides has been investigated in aqueous solutions of sulfuric, perchloric, and hydrochloric acid at 85.0 ± 0.05 °C. Analysis of the kinetic data by the excess acidity method, Arrhenius parameters, the order of the catalytic effects of strong acids, the kinetic deuterium isotope effect, and the substituent effect have indicated that the hydrolysis of 5‐substituted benzosulfamides 1a , 1b , 1c , 1d occur with a mechanistic switchover from A2 to A1 in the studied range: an A2 mechanism in low acidity regions and an A1 mechanism in high acid concentrations. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

2.
Disproportionation of cyclic nitroxyl radicals (NRs) in acid solutions is of key importance for the chemistry of these compounds. Meanwhile, the data reported on the mechanism of this reaction in dilute acids are inconsistent with those on the stability of NRs in concentrated acids. Here we have examined the kinetics and stoichiometry for the disproportionation of 2,2,6,6‐tetramethylpiperidine‐1‐oxyl ( 1 ) in aqueous H2SO4 (1.0–99.3 wt%) and found that (1) the disproportionation of 1 proceeds by the same mechanism over the entire range of acid concentrations, (2) the effective rate constant of the process exhibits a bell‐shaped dependence on the excess acidity function X peaked at X = ?pK 1H+ = 5.8 ± 0.3, (3) a key step of the process involves the oxidation of 1 with its protonated counterpart 1H + yielding oxopiperidinium cation 2 and hydroxypiperidine 3 at a rate constant of (1.4 ± 0.8) × 105 M?1 · s?1, and (4) the reaction is reversible and, upon neutralization of acid, disproportionation products 2 and 3H + comproportionate to starting 1 . In highly acidic media, the protonated form 1H + is relatively stable due to a low disproportionation rate. Based on the known and newly obtained values of equilibrium constants, both the standard redox potential for the 1H + / 3 pair (955 ± 15 mV) and the pH‐dependences have been calculated for the reduction potentials of 1 and 2 to hydroxylamine 3 that is in equilibrium with its protonated 3H + and deprotonated 3 ? forms. The data obtained provide a deeper insight into the mechanism of nitroxyl‐involving reactions in chemical and biological systems. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

3.
Theoretical computations and experimental kinetic measurements were applied in studying the mechanistic pathways for the alkaline hydrolysis of three secondary amides: N‐methylbenzamide, N‐methylacetamide, and acetanilide. Electronic structure methods at the HF/6‐31+G(d,p) and B3LYP/6‐31+G(d,p) levels of theory are employed. The energies of the stationary points along the reaction coordinate were further refined via single point computations at the MP2/6‐31+G(d,p) and MP2/6‐311++G(2d,2p) levels of theory. The role of water in the reaction mechanisms is examined. The theoretical results show that in the cases of N‐methylbenzamide and N‐methylacetamide the process is catalyzed by an ancillary water molecule. The influence of water is further assessed by predicting its role as bulk solvent. The alkaline hydrolysis process in aqueous solution is characterized by two distinct free energy barriers: the formation of a tetrahedral adduct and its breaking to products. The results show that the rate‐determining stage of the process is associated with the second transition state. The entropy terms evaluated from theoretical computations referring to gas‐phase processes are significantly overestimated. The activation barriers for the alkaline hydrolysis of N‐methylbenzamide and acetanilide were experimentally determined. Quite satisfactory agreement between experimental values and computed activation enthalpies was obtained. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

4.
Electronic structure calculations and second‐order delocalizations in N,N′,N′′‐triaminoguanidine ( TAG ) have been studied by employing ab initio MO and density functional methods. There are total 10 rotational isomers on the potential energy (PE) surface of TAG . The effect of three amino groups substitution on guanidine ( Gu ) has been studied in terms of the primary and the secondary electron delocalizations in TAG by employing Natural Population Analysis (NPA). An increased electron delocalization is observed in protonated triaminoguanidine ( TAGP ) due to the three strong intramolecular hydrogen bonds and hence accounts for its extra stability. The increase in the electron delocalization upon protonation in TAG can be compared to that in guanidine. The absolute proton affinity (APA) of TAG is less than that of Gu . HOMA and NICS studies have been carried out to understand electron delocalization in TAGP . Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

5.
In the present study, we explore the application of several theoretically estimated indices that characterize the reactivity of a series of phenyl N‐phenylcarbamates in the alkaline hydrolysis reaction. The rate constants (at 25 °C) for the hydrolysis of several derivatives were spectrophotometrically determined. The obtained kinetic data in this study, combined with literature data for other derivatives, were then correlated with theoretically estimated reactivity indices: Hirshfeld and NBO atomic charges, the Parr electrophilicity index (ω), and the electrostatic potential at the carbon and oxygen atoms of the reaction centre (VC, VO). The predictive ability of these quantities is discussed in a comparative context. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

6.
The reactivity of 6‐quinolinyl and 8‐quinolinyl N,N‐dimethylcarbamates was examined in several aqueous basic media. A quadratic dependence was observed for the constant rates upon hydroxide concentration for both compounds, which is a typical behaviour of a mechanism involving a base‐catalysed deprotonation of the tetrahedral intermediate with the formation of a dianion at high concentrations of hydroxide ion, while at lower concentrations a specific‐base catalysed addition–elimination mechanism seems to be predominant. The reactivity of 8‐quinolinyl N,N‐dimethylcarbamate was also studied in several amine buffers, showing specific base catalysis. The reactivity of 6‐quinolinyl N,N‐dimethylcarbamate was studied in H2O and in D2O and the solvent isotope effect supports the proposal of a mechanism involving a specific‐base hydrolysis. All results confirm the existence of a mechanism with a rate determining step involving the substrate anion and a second mole of hydroxide ion. This mechanism was so far unknown for carbamate reactivity, being only known to occur with amides. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

7.
Infrared (IR) and Raman spectra were obtained for N,N′‐dicyclohexylcarbodiimide (DCC) in the solid state and in CHCl3 solution. Structures and vibrational spectra of isolated, gas‐phase DCC molecules with C2 and Ci symmetries, computed at the B3‐LYP/cc‐pVTZ level, show that the IR and Raman spectra provide convincing evidence for a C2 structure in both the solid state and in CHCl3 solution. Using a scaled quantum‐chemical force field, these density functional theory calculations have provided detailed assignments of the observed IR and Raman bands in terms of potential energy distributions. Comparison of solid‐state and solution spectra, together with a Raman study of the melting behaviour of DCC, revealed that no solid‐state effects were evident in the spectra. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

8.
In the current work, density functional theory calculations were performed to elucidate the detailed reaction mechanism for N‐heterocyclic carbene (NHC)‐catalyzed oxidative N‐acylation reaction of amides with aldehydes affording imide products. According to the calculated results, the reaction is initiated by the nucleophilic attack of NHC to aldehydes forming zwitterionic intermediate, which can then form Breslow intermediate via proton transfer. The Breslow intermediate can then be oxidized affording the oxidative intermediate, which can then go through 1,2‐addition with the deprotonated N‐sulfonylcarboxamides. Subsequently, elimination of NHC catalyst produces the final imide product. Our results reveal that the proton in N‐sulfonylcarboxamides is probably abstracted by base t‐BuOK or DPQH, and the deprotonation process is barrier‐less. Moreover, for the second step, ie, the formation of Breslow intermediate, direct proton transfer is impossible to occur. On the contrary, the results reveal that t‐BuOH can mediate the proton transfer in this step and significantly lower the energy barrier to 24.1 kcal/mol, which is also the highest energy barrier for the whole reaction. The work provides not only valuable clues for elucidating the detailed reaction mechanism for the invaluable NHC‐catalyzed oxidative reactions but also mechanistic insights for the rational design of novel NHC‐catalyzed oxidative reactions in the future.  相似文献   

9.
The gas‐phase elimination kinetics of 2,2‐diethoxyethyl amine and 2,2‐diethoxy‐N,N‐diethylethanamine (320–380 °C; 40–150 Torr) in a seasoned reaction vessel are homogeneous, unimolecular and obey a first‐order rate law. These elimination processes involve two parallel reactions. The first gives ethanol and the corresponding 2‐ethoxyethenamine. The latter compound further decomposes to ethylene, CO and the corresponding amine. The second parallel reaction produce ethane and the corresponding ethyl ester of an α‐amino acid. The following Arrhenius expressions are given as: For 2,2‐diethoxyethyl amine For 2,2‐diethoxy‐N,N‐diethylethanamine Comparative kinetic and thermodynamic parameters of the overall, the parallel and the consecutive reactions lead to consider two types of mechanisms in terms of a concerted polar cyclic transition state structures. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

10.
Some initial acid‐catalytic reactions involved in the synthesis of the urea‐formaldehyde resin were theoretically investigated at B3LYP and MP2 levels with solvent effects included. The results suggest that the addition between urea and formaldehyde in neutral condition undergoes with a concerted mechanism represented by a four‐member ring transition state. For this reaction, a notable barrier (above 130 kJ/mol) was identified at all theoretical levels. The reactions between urea and different protonated forms of formaldehyde in acid solution were investigated. The reaction between protonated methanediol with urea can produce the methylol urea cation via an SN2 transition state with a lower barrier of 54.8 kJ/mol. With the mediation of a water molecule, the intra‐molecular proton transfer produces the stable methylol carbonium (NH2CONHCH2+), which plays an important role in the following formation of methylene and methylene ether linkages. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

11.
We have demonstrated a simple near‐infrared (NIR) photodetector (PD) based on charge transfer complex (CTC) formed in molybdenum trioxide (MoO3) doped N,N′‐di(naphthalene‐1‐yl)‐N,N′‐diphenyl‐benzidine (NPB), which shows a photocurrent of about 0.35 A/cm2 at –3 V under 980 nm illumination. The existence of CTC formation promotes photocurrent generation which is investigated by comparison with MoO3 doped 2‐methyl‐9,10‐di(2‐naphthyl)anthracene (MADN) film which has no CTC absorption. It can be evolved that this kind of simple‐structure photodetector has potential application in the near‐infrared (NIR) detection area. It is shown in this Letter that although both MoO3 and NPB have larger energy gaps of about 3 eV and weak absorption in the NIR region, the charge transfer complexes formed by mixing the two materials show an extra absorption band and good photoelectric response in the NIR region.

  相似文献   


12.
The conversion of the Z‐phenylhydrazone of 5‐amino‐3‐benzoyl‐1,2,4‐oxadiazole ( 1a ) into the relevant 1,2,3‐triazole ( 2a) has been quantitatively studied in toluene in the presence of several halogenoacetic acids ( HAA s, 3a – h ). Again, the occurrence of two reaction pathways has been pointed out: they require one or two moles of acid, respectively, thus repeating the situation previously observed in the presence of trichloroacetic acid. The observed rate constant ratios (kIII/kII) are only slightly affected by the nature of the acid used. To gain a deeper insight into the action of the acids used we have measured the association constants of the HAA s ( 3a – h) with 4‐nitroaniline ( 4 ) in toluene. Also in this case, the formation of two complexes requiring one (K2) or two (K3) moles of acid has been evidenced, but now the K3/K2 ratios are significantly affected by the strength of the acid examined. The variation of the K3/K2 ratios larger than those concerning the kIII/kII ratios appears useful to enlighten the very nature of the acid‐catalyzed pathways in toluene, which has been elucidated also carrying out the rearrangement in the presence of mixtures of tribromo‐ and trichloro‐acetic acids at different concentrations. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

13.
The kinetics of hydrolysis of 1,8‐N‐butyl‐naphthalimide (1,8‐NBN) to 1,8‐N‐butyl‐naphthalamide (1,8‐NBAmide) and of 2,3‐N‐butyl‐naphthalimide (2,3‐NBN) to 2,3‐N‐butyl‐naphthalamide (2,3‐NBAmide), as well as the formation of the respective anhydrides from the amides were investigated in a wide acidity range. 1,8‐NBN equilibrates with 1,8‐NBAmide in mild alkali. Under the same conditions 2,3‐NBN quantitatively yields 2,3‐NBAmide. Over a wide range of acidities the reactions of the 1,8‐ and 2,3‐N‐butyl‐naphthalamides (or imides) yield similar products but with widely different rates and at distinct pH's. Anhydride formation in acid was demonstrated for 1,8‐NBAmide. The reactions mechanisms were rationalized in the manifold pathways of ab initio calculations. The differences in rates and pH ranges in the reactions of the 1,8‐ and 2,3‐N‐butyl‐naphthalamides were attributed to differences in the stability of the tetrahedral intermediates in alkali as well as the relative stabilities of the five and six‐membered ring intermediates. The rate of carboxylic acid assisted 1,8‐N‐Butyl‐naphthalamide hydrolysis is one of the largest described for amide hydrolysis models. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

14.
15.
The intramolecular asymmetric Michael addition reaction catalyzed by imidazolidinone is investigated using the density functional theory calculations. The details of the reaction mechanism, potential energy surfaces, and the influence of the acid additive are investigated. The reaction process includes two stages. The first stage is Michael addition, in which the enamine complex is created and then the Michael addition is carried out. The second stage is a product separation stage which includes an enol‐keto tautomerization and a two‐step hydrolysis. The enantioselectivity is controlled by the Michael addition step which involves a new carbon–carbon bond formation. The calculation results provide a general model which may explain the mechanism and enantioselectivity of the title reaction. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

16.
The mechanisms of the single and double Mannich reactions between acetaldehyde and N‐Boc imines are clarified by density functional theory calculations. For single addition of Mannich reaction, the energy difference between the transition states of different configurations correspond to an enantiomeric excess value of 90.58% (without solvent) and 98.46% (in acetonitrile) in favor of the (S)‐configuration product. For bis‐addition of Mannich reaction, the calculated enantiomeric excess value is 95.02% (without solvent) and 98.57% (in acetonitrile) in favor of the (S, S)‐configuration product. These calculated results are in good agreement with the experimental results. The calculations clearly demonstrate that the hydrogen‐bonding determine the stereochemistry of the reactions. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

17.
We have isolated two isomeric solids 1 and 2 of N,N′‐bis(3,5‐dichlorosalicylidene)‐2,2′‐ethylenedianiline and characterized by IR, UV/Vis, X‐ray powder diffraction, thermogravimetric analysis/differential thermal analysis, and X‐ray crystallography. Although the solids are same formulas, each shows different colors and crystal structures. Orange solid ( 1 ) shows endo conformation while yellow solid ( 2 ) exhibits exo form depending on packing modes. UV/Vis spectra of 1 and 2 appear very similar patterns in the solid state; however, the bands of 1 are slightly red‐shifted compared with those of 2 . 1 displays a strong fluorescent emission band at ~582 nm while 2 shows an intense fluorescent signal at ~563 nm. The charge density populations of 1 and 2 have been studied by computational simulations using density functional theory at pbe1pbe/6‐311G** level. The calculated highest occupied molecular orbital and lowest unoccupied molecular orbital energies of 1 and 2 confirm that charge transfer occurs within the organic molecules. The energy difference of HOMO‐LUMO in 1 is smaller slightly than that of 2 about 0.05 eV (~17 nm). Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

18.
Acid‐catalyzed disproportionation of cyclic nitroxyl radicals R2NO? includes the half‐reactions of their oxidation to oxoammonium cations R2NO+ and reduction to hydroxylamines R2NOH. For many nitroxyl radicals, this reaction is characterized by its ~100% reversibility. Quantitative characteristics of acid–base and redox properties of the whole redox triad may be obtained from research of kinetics and equilibrium of this reaction. Here, we have examined the kinetics for the disproportionation of twenty piperidine‐, pyrroline‐, pyrrolidine‐, and imidazoline nitroxyl radicals in aqueous H2SO4, and interpreted it in terms of the excess acidity function X. The rate‐limiting step of this reaction is R2NO? oxidation by its protonated counterpart R2NOH+?. Kinetic stability of R2NO? in acidic media depends on the basicity of nitroxyl group. This basicity is influenced predominantly by protonation of another, more basic group in radical structure, and its proximity to nitroxyl group. The discovered estimates of pK values for radical cations R2NOH+? (from ?5.8 to ?12.0) indicate a very low basicity of nitroxyl groups in all commonly used R2NO?. For the first time, a linear correlation is obtained between the one‐electron reduction potentials of oxoammonium cations and the basicity of nitroxyl groups of related radicals. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

19.
The kinetics of base catalyzed cyclization of 2,6‐dinitrophenylsulfanyl ethanenitrile and 2,4,6‐trinitrophenylsulfanyl ethanenitrile giving 2‐cyano‐7‐nitrobenzo[d]thiazole‐3‐oxide and 2‐cyano‐5,7‐dinitrobenzo[d]thiazole‐3‐oxide respectively was studied in methanolic methoxyacetate, acetate, trichlorophenoxide, N‐methylmorpholine, and N‐methylpiperidine buffers at 25 °C and I = 0.1 mol L?1. It was found that reaction involves both general acid and general base catalyses whose manifestation depends on the pKa of the acid‐buffer component and the ratio of both buffer components. In weakly basic buffers the rate‐limiting step is C? H bond breaking in the cyclic intermediate, while in strongly basic buffers the rate‐limiting step is the general acid‐catalyzed elimination of hydroxyl group from the intermediate. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

20.
Recent studies have shown that general‐base assisted catalysis is a viable mechanistic pathway for hydrolysis of smaller anhydrides. Therefore, it is the central purpose of the present work to compare and contrast the number of hydrogen atoms in‐flight and stationary in the transition state structure of the base‐catalyzed mechanisms of 2 hydrolytic reactions as well as determine if any solvent effects occur on the mechanisms. The present research focuses on the hydrolytic mechanisms of N,N‐dimethylformamide (DMF) and acetic anhydride in alkali media of varying deuterium oxide mole fractions. Acetic anhydride has been included in this study to enable comparisons with DMF hydrolysis. Comparative studies may give synergistic insight into the detailed structural features of the activated complexes for both systems. Hydrolysis reactions in varying deuterium oxide mole fractions were conducted in concentrations of 2.0M , 2.5M , and 3.0M for DMF and 0.10M for acetic anhydride at 25°C. Studies in varying deuterium mole fractions allow for proton inventory analysis, which sheds light on the number and types of hydrogen atoms involved in the activated complex. For these systems, this type of study can distinguish between direct nucleophilic attack of the hydroxide ion on the carbonyl center and general‐base catalysis by the hydroxide ion to facilitate a water molecule attacking the carbonyl center. The numerical data are used to discuss 3 possible mechanisms in the hydrolysis of DMF.  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