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 31 毫秒
1.
The polymerization of acrylonitrile (AN) initiated by tetramethyl-2-tetrazene (TMT) and p-toluenesulfonic acid (TSA) in dimethylformamide (DMF) was studied. The polymerization was confirmed to proceed through a radical mechanism. The initial rate of polymerization R was expressed by the equation: Rp = k[TMT]0.6 [TSA]0.46 [AN]1.35. The overall activation energy for the polymerization was estimated as 20.7 kcal/mole. In the absence of monomer, the reaction of TMT with TSA was also studied kinetically by measuring the evolution of nitrogen. From these results and ESR measurement of the TMT/TSA system, a possible initiation mechanism is proposed.  相似文献   

2.
The polymerization of acrylonitrile (AN) initiated by the system of tetramethyl tetrazene (TMT) and bromoacetic acid (BA) in dimethylformamide (DMF) was studied. The TMT–BA system could initiate the polymerization of AN more easily than TMT alone. The polymerization was confirmed to proceed through a radical mechanism. The initial rate of polymerization Rp was expressed by the equation: Rp = [TMT]0.62-[BA]0.5[AN]1.5. The overall activation energy for the polymerization was estimated as 9.4 kcal/mole. In the absence of monomer, the reaction of TMT with BA in DMF was also studied kinetically by measuring the evolution of nitrogen gas. The reaction was first-order in TMT and first-order in BA; the rate data at 49°C were k2 = 9.1 × 10?2l./mole-sec., ΔH? = 17.0 kcal/mole, and ΔS? = ? 6.6 eu. In addition, the treatment of TMT with BA in benzene led to the formation of tetramethylhydrazine radical cation, which was identified by its ESR spectrum. On the other hand, the relatively strong interaction between TMT and DMF was observed by absorption spectrophotometry.  相似文献   

3.
The interactions of a weak organic acid (acetic acid, HOAc) with a toluene solution of the Co(II)-Schiff base type complex, (R,R')-N,N'-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexane-diamino Co(II) (labeled [Co(1)]), was investigated using EPR, HYSCORE, and DFT computations. This activated [Co(II)(1)] system is extremely important within the context of asymmetric catalysts (notably the hydrolytic kinetic resolution of epoxides) despite the lack of detailed structural information about the nature of the paramagnetic species present. Under anaerobic conditions, the LS [Co(II)(1)] complex with a |yz, (2)A(2)? ground state is converted into a low-spin (LS) and a high-spin (HS) complex in the presence of the acid. The newly formed LS state is assigned to the coordinated [Co(II)(1)]-(HOAc) complex, possessing a |z(2), (2)A(1)? ground state (species A; g(x) = 2.42, g(y) = 2.28, g(z) = 2.02, A(x) = 100, A(y) = 120, A(z) = 310 MHz). The newly formed HS state is assigned to an acetate coordinated [Co(II)(1)]-(OAc(-)) complex, possessing an S = (3)/(2) spin ground state (species B, responsible for a broad EPR signal with g ≈ 4.6). These spin ground states were confirmed with DFT calculations using the hybrid BP86 and B3LYP functionals. Under aerobic conditions, the LS and HS complexes (species A and B) are not observed; instead, a new HS complex (species C) is formed. This complex is tentatively assigned to a paramagnetic superoxo bridged dimer (AcO(-))[Co(II)(1)···O(2)(-)Co(III)(1)](HOAc), as distinct from the more common diamagnetic peroxo bridged dimers. Species C is characterized by a very broad HS EPR signal (g(x) = 5.1, g(y) = 3.9, g(z) = 2.1) and is reversibly formed by oxygenation of the LS [Co(II)(1)]-(HOAc) complex to the superoxo complex [Co(III)(1)O(2)(-)](HOAc), which subsequently forms the association complex C by interaction with the HS [Co(II)(1)](OAc(-)) species. The LS and HS complexes were also identified using other organic acids (benzoic and propanoic acid). Thermal annealing-quenching experiments revealed the additional presence of [Co(III)(1)O(2)(-)](HOAc) adducts, corroborating the presence of species C and the presence of diamagnetic dimer complexes in the solution, such as the EPR silent (HOAc)[Co(III)(1)(O(2)(2-))Co(III)(1)](HOAc). Overall, it appears that a facile interconversion of the [Co(1)] complex, possessing a LS ground state, occurs in the presence of acetic acid, producing both HS and LS Co(II) states, prior to formation of the oxidized active form of the catalyst, [Co(III)(1)](OAc(-)).  相似文献   

4.
The oxidation of hydroquinone by environmentally benign tetrabutyl ammonium tribromide (TBATB) was carried out in 50% V/V aqueous acetic acid medium under pseudo-first-order conditions, keeping a large excess of hydroquinone over the oxidant. The main reactive species of oxidant and substrate were found to be the Br3-\mathrm{Br}_{3}^{-} ion and hydroquinone, respectively. The reaction proceeds with prior complex formation between the reactants followed by its slow decomposition to generate semiquinone and bromine radicals. The complex formation was kinetically verified by its Michaelis–Menten plot. The solvent effect was verified by using Grunwald–Winstein equation which is consistent with an SN2 mechanism. The formation constants for the complex and rate constant for the slow decomposition step were determined by studying the reaction at five different temperatures. The values of formation constant of the complex and the rate constant for its decomposition were determined at these temperatures. The activation parameters with respect to the slow step of the reaction have also been determined.  相似文献   

5.
The excited-state proton transfer and subsequent intramolecular ion pair formation of a cupreidine-derived Cinchona organocatalyst () were studied in THF-water mixtures using picosecond time-resolved fluorescence together with global analysis. Full spectral and kinetic characterization of all the fluorescent species allowed us to monitor the 3-step process for the ion pair dissociation. In the first step, proton transfer occurs through a water "wire" from the 6-hydroxyquinoline unit (excited-state acid) to the covalently bonded basic quinuclidine moiety, resulting in a hydrogen bonded ion pair. This was confirmed by the observed kinetic isotope effect in the presence of heavy water. In the second step, the formed ions are further solvated by a few solvent molecules, producing the solvent separated ion pair. Finally, a fully solvated ion pair is formed. The 5-exponential global model derived from the reaction scheme describes the experimental data very well.  相似文献   

6.
以石油焦基微晶碳作为电极材料,并由N2吸附,X射线衍射(XRD)表征其孔结构和微晶结构.研究了4种电解液Et4NBF4/PC(四乙基铵四氟硼酸盐/碳酸丙烯酯)、Et4NBF4/AN(四乙基铵四氟硼酸盐/乙腈)、Bu4NBF4/PC(四丁基铵四氟硼酸盐/碳酸丙烯酯)和Bu4NBF4/AN(四丁基铵四氟硼酸盐/乙腈)的微晶碳电容器特性.结果表明:电解质离子与溶剂AN的溶剂化半径较小,容易嵌入类石墨微晶碳层,其于AN的电活化电位比在PC中的低,致使电活化程度更深,材料的表面利用率更高,电容量较大.电解质阳离子(Et4N+,Bu4N+)大小对电活化影响不大.电活化使材料类石墨微晶层间距(d002)变大,离子尺寸越大,层间距增加越明显.  相似文献   

7.
Data on the special salt effect in monomolecular heterolysis reactions (Sn1, E1, solvolysis) are summarized and critically analyzed. The mechanisms suggested by Ingold, Winstein, Dannenberg, Okamoto, and the authors are discussed. The special salt effect is due to the effect of a salt on the contact ion pair of a substrate. Quadrupoles and ion triplets are formed. In the limiting step of the heterolysis, a contact ion pair interacts with a solvent cavity. Association of salts with a contact ion pair increases the lifetime of the cationoid and the probability of its contact with the solvent cavity. A spatially separated ion pair is formed, which rapidly transforms into a solvation-separated ion pair, which, also rapidly, yields reaction products.  相似文献   

8.
To understand the detailed mechanism of a recently reported tandem iminium cyclization and Smiles rearrangement, the reaction processes of a chiral substrate were investigated by monitoring its stereochemical courses. Under the tandem reaction conditions, chiral aldehyde 1 derived from l-prolinol led to two surprising results. First, the iminium cyclization gave a diastereomeric mixture with the cis-configured product as the predominant one. Second, Smiles rearrangement of both cis- and trans-2 led to the same product 3a directly derived from the trans isomer. The former was rationalized by the postulation of a Cram's chelate transition state leading to the cis product as kinetically favored. The latter was due to the equilibration between the trans/cis pair involving a carbocation intermediate and the steric hindrance, which prevented the cis isomer from undergoing the intramolecular nucleophilic substitution. This hypothesis was further supported by the results of a competition experiment in which the addition of 1 equiv of p-methoxyaniline in the rearrangement step led to a significant amount of anilinyl-exchanged rearrangement product.  相似文献   

9.
The acetylation of tert-butanol with acetic anhydride catalyzed by 4-(dimethylamino)pyridine (DMAP) has been studied at the Becke3 LYP/6-311 + G(d,p)//Becke3 LYP/6-31G(d) level of theory. Solvent effects have been estimated through single-point calculations with the PCM/UAHF solvation model. The energetically most favorable pathway proceeds through nucleophilic attack of DMAP at the anhydride carbonyl group and subsequent formation of the corresponding acetylpyridinium/acetate ion pair. Reaction of this ion pair with the alcohol substrate yields the final product, tert-butylacetate. The competing base-catalyzed reaction pathway can either proceed in a concerted or in a stepwise manner. In both cases the reaction barrier far exceeds that of the nucleophilic catalysis mechanism. The reaction mechanism has also been studied experimentally in dichloromethane through analysis of the reaction kinetics for the acetylation of cyclohexanol with acetic anhydride, in the presence of DMAP as catalyst and triethylamine as the auxiliary base. The reaction is found to be first-order with respect to acetic anhydride, cyclohexanol, and DMAP, and zero-order with respect to triethyl amine. Both the theoretical as well as the experimental studies strongly support the nucleophilic catalysis pathway.  相似文献   

10.
The kinetics of polymerization of acrylonitrile(AN) initiated by manganese(III) acetate in the presence of glycerol was investigated in the temperature range of 30–40°C. The effect of varying the concentrations of glycerol, sulfuric acid, acetic acid, metal ion, and monomer on the rate was studied. A suitable reaction scheme and rate expression have been proposed. Termination was mutual and was caused by the combination of two growing polymer radicals.  相似文献   

11.
The photochemical reactions of a series of triaryl vinyl halides 1X in acetic acid and in acetonitrile have been studied using product analysis as a function of the time of irradiation. The quantum efficiencies of formation of the products derived from the photogenerated vinyl cations 1(+) depend on the alpha-aryl substituent, the beta-aryl substituent, the leaving group X (= bromide or chloride), and the temperature at which the irradiations are carried out. Hammett correlation or noncorrelation of the alpha-aryl substituent effects with (excited-state) substituent constants indicates that the ions 1(+) are formed directly from the excited states of 1X by heterolytic cleavage of the carbon-halogen bond. Homolytic cleavage, yielding radicals 1(*), is a parallel process: the partitioning into ion and radical occurs in the excited state. This conclusion is corroborated by the leaving group effect and the temperature effect. By performing the photochemical reactions of 1X in the presence of HOAc and/or NaOAc as well as the labeled common halide ion (82)Br(-) or (36)Cl(-), the relative reactivities of the cations 1(+) toward these nucleophiles were determined. The selectivities follow the Reactivity-Selectivity Principle. The temperature effect data show that the reactions of the cations with the anionic nucleophiles are (de)solvation controlled and their reactions with the neutral nucleophile activation controlled.  相似文献   

12.
Observed pseudo-first-order rate constants (k(obs)) of the hydride-transfer reactions from isopropyl alcohol (i-PrOH) to two NAD(+) analogues, 9-phenylxanthylium ion (PhXn(+)) and 10-methylacridinium ion (MA(+)), were determined at temperatures ranging from 49 to 82 degrees C in i-PrOH containing various amounts of AN or water. Formations of the alcohol-cation ether adducts (ROPr-i) were observed as side equilibria. The equilibrium constants for the conversion of PhXn(+) to PhXnOPr-i in i-PrOH/AN (v/v = 1) were determined, and the equilibrium isotope effect (EIE = K(i-PrOH)/K(i-PrOD)) at 62 degrees C was calculated to be 2.67. The k(H) of the hydride-transfer step for both reactions were calculated on the basis of the k(obs) and K. The corresponding deuterium kinetic isotope effects (e.g., KIE(OD)(H) = k(H)(i-PrOH)/k(H)(i-PrOD) and KIE(beta-D6)(H) = k(obs)(i-PrOH)/k(obs)((CD3)2CHOH)), as well as the activation parameters, were derived. For the reaction of PhXn(+) (62 degrees C) and MA(+) (67 degrees C), primary KIE(alpha-D)(H) (4.4 and 2.1, respectively) as well as secondary KIE(OD)(H) (1.07 and 1.18) and KIE(beta-D6)(H) (1.1 and 1.5) were observed. The observed EIE and KIE(OD)(H) were explained in terms of the fractionation factors for deuterium between OH and OH(+)(OH(delta+)) sites. The observed inverse kinetic solvent isotope effect for the reaction of PhXn(+) (k(obs)(i-PrOH)/k(obs)(i-PrOD) = 0.39) is consistent with the intermolecular hydride-transfer mechanism. The dramatic reduction of the reaction rate for MA(+), when the water or i-PrOH cosolvent was replaced by AN, suggests that the hydride-transfer T.S. is stabilized by H-bonding between O of the solvent OH and the substrate alcohol OH(delta+). This result suggests an H-bonding stabilization effect on the T.S. of the alcohol dehydrogenase reactions.  相似文献   

13.
Mechanistic studies on the B(C(6)F(5))(3) catalyzed allylstannation of isomeric substituted benzaldehydes are reported. Confirming a report by Maruoka et al., good (5:1) to excellent (>20:1) selectivities for ortho over para isomers are observed when 1:1 mixtures (X = OMe, Cl, F, OTBS) are allylstannated with C(3)H(5)SnBu(3) in the presence of B(C(6)F(5))(3) (2.5% per CHO). The best selectivities are observed for the anisaldehydes. Multinuclear NMR studies on solutions of B(C(6)F(5))(3) and C(3)H(5)SnBu(3) (1:1 to 1:5) show that the borane abstracts the allyl group from the organotin reagent, forming an adduct (C(6)F(5))(3)B...CH(2)CHCH(2)SnBu(3), 1, or ion pair [(C(6)F(5))(3)BCH(2)CH=CH(2)](-)[Bu(3)SnCH(2)CHCH(2)SnBu(3)](+), 2, depending on the reagent ratio. These compounds are important in the mechanism of Lewis acid catalyzed 1,3-isomerization of substituted allyl stannanes. When allyltin reagent is added to solutions of B(C(6)F(5))(3) and ortho-anisaldehyde (1:5) at -60 degrees C, conversion to the stannylium ion pair [Bu(3)Sn(ortho-anisaldehyde)(2)](+)[o-ArCH(allyl)OB(C(6)F(5))(3)](-), o,o-4, is observed. The structure of this species was confirmed by (1)H, (11)B, (19)F, and (119)Sn NMR spectroscopy and by forming related ion pairs (o-5 and o,o-5) utilizing the [B(C(6)F(5))(4)](-) counteranion via reaction of [Bu(3)Sn](+)[B(C(6)F(5))(4)](-) with aldehyde. The anion in o,o-4 is formed via direct allylation of the ortho-anisaldehyde/B(C(6)F(5))(3) adduct o-3, while the cation arises upon aldehyde ligation of the resulting tributylstannylium ion. The crystal structure of the related derivative ortho-C(6)H(4)(OMe)CHO x SnMe(3)BF(4), 6, showed that the aldehyde binds the tin nucleus only through the carbonyl oxygen. Similar reactions using para-anisaldehyde show that formation of p,p-4 occurs at a much slower rate, again demonstrating the preference for the ortho substituted substrates. For similar experiments using benzophenone, however, formation of the ion pair [Bu(3)Sn(Ph(2)CO)(2)](+)[(C(3)H(5))B(C(6)F(5))(3)](-), 8, was observed, illustrating the differences subtle changes in substrate can bring. Ion pair 8 is formed via the trapping of 1 by the benzophenone substrate. In the presence of excess aldehyde and allyltin reagent, ion pair o,o-4 catalyzes the allylstannation of aldehyde to give the product stannyl ether. Several lines of experimental evidence suggest this is the true catalyst in the system. The chemoselectivity observed thus does not rely on classical chelation control in any way. Rather, we propose that the ortho donor group stabilizes the developing positive charge at the beta carbon of the allyl group and the tin atom during the allylation event. This stabilization renders the ortho substituted substrates kinetically favored toward allylation irrespective of the Lewis acid employed.  相似文献   

14.
Multiple kinetic isotope effects (KIEs) on deoxyadenosine monophosphate (dAMP) hydrolysis in 0.1 M HCl were used to determine the transition state (TS) structure and probe its intrinsic reactivity. The experimental KIEs revealed a stepwise (SN1) mechanism, with a discrete oxacarbenium ion intermediate. This is the first direct evidence for the deoxyribosyl oxacarbenium ion in solution. In 50% methanol/0.1 M HCl the products were deoxyribose 5-phosphate (dRMP) and alpha- and beta-methyl dRMP. The alpha-Me-dRMP/beta-Me-dRMP ratio was 8.5:1. Assuming that a free oxacarbenium ion is equally susceptible to nucleophilic attack on either face, this indicated that approximately 20% proceeded through a solvent-separated ion pair complex, or free oxacarbenium ion, a DN+AN mechanism, while approximately 80% of the reaction proceeded through a contact ion pair complex. The oxacarbenium ion lifetime was estimated at 10(-11)-10(-10) s. Computational transition states were found for ANDN, DN*AN, DN*AN, and DN+AN mechanisms using hybrid density functional theory calculations. After taking into account 20% of DN+AN, there was an excellent match of calculated to experimental KIEs for 80% of the reaction having a DN*AN mechanism. That is, C-N bond cleavage is reversible, with dAMP and the {oxacarbenium ion*adenine} complex in equilibrium. The first irreversible step is water attack on the oxacarbenium ion. The calculated 1'-14C KIE for a stepwise mechanism with irreversible C-N bond cleavage (DN*AN) was 1.052, in the range previously associated only with ANDN transition states, and close to the calculated ANDN value, 1.059. The 1'-14C KIE was strongly dependent on the adenine protonation state.  相似文献   

15.
Dihaloindium hydrides (X2InH) are novel reducing reagents, which act in both an ionic and a radical manner. The hydrides were easily generated from InX3 and Bu3SnH to reduce a variety of functionalities such as aldehydes, ketones, enones, and imines. The combination of a phosphine and Cl2InH accomplished the selective transformation from acid chlorides to aldehydes. One-pot treatment of Cl2InH, enones, and aldehydes achieved reductive aldol reactions, in which the predominant reduction of enones was followed by an aldol reaction between the resulting indium enolates and the remaining aldehydes. It is noteworthy that both anti- and syn-selective aldols were obtained by the use of THF and an aqueous solvent, respectively. The replacement of Bu3SnH with Et3SiH as a hydride source allowed the catalytic use of InBr3 to give the syn-selective aldols. The dehalogenation of alkyl halides was achieved by a catalytic amount of InCl3 in the presence of Bu3SnH. This procedure was applied to some representative cyclizations as radical proof. A simple and non-toxic system, NaBH4/InCl3, also promoted dehalogenation, intramolecular cyclization, and intermolecular coupling reactions. In addition, the Et3SiH/InCl3 system was found applicable to an effective intramolecular cyclization of enynes.  相似文献   

16.
The hybrid imine/amine palladium(II) coordination complexes [PdX2(kappa2-N(imino),N(amino))](X = Cl, AcO; kappa2-N(imino),N(amino)= 4ClC6H4CHNCH2(CH2)nN(CH3)2, n= 1, 2) have been prepared in different isomeric forms which include E/Z arrangement around the C[double bond]N bond of the hybrid ligand and {Pd(kappa(2)-N(imino),N(amino))} ring conformation. The crystal structures of four of them, E-1AcO, Z-1AcO, E-2AcO and E-2Cl, have been determined and the solution behaviour in acetic acid, the common cyclometallating solvent, for all these systems studied. The complexes in acetic acid solution are shown to maintain the structure determined by X-ray crystallography, as they do in deuterated chloroform. Nevertheless, a partial opening equilibrium of the {Pd(kappa2-N(imino),N(amino))} ring is observed by NMR experiments. When the complexes are held in solution for longer periods the corresponding cyclometallated derivatives, 1AcO-CM, 2AcO-CM, 1Cl-CM and 2Cl-CM, containing the {Pd(kappa2-C,N(imino))} palladacycle are obtained, as characterized by 1H NMR spectroscopy. In these compounds the total opening of the N(amino) moiety of the ligand has occurred. The C-H bond activation process has been studied kinetico-mechanistically at different temperatures, pressures and acid concentrations; the results agree with the need of an opening of the chelate ring in [PdX2(kappa2-N(imino),N(amino))] prior to the proper cyclometallation reaction. The values of the enthalpies of activation are higher than those observed for known N-monodentated cyclometallating ligands, as should correspond to the contribution of a ligand dechelation pre-equilibrium. The entropies and volumes of activation are also indicative of this predissociation that include an important amount of contractive ordering. The presence of small amounts of triflic acid in the reaction medium accelerates the reaction to the value observed for N(imino)-monodentate systems, indicating that the full opening of the chelate ring has taken place. For the badly oriented isomeric forms of the ligand in the chelated complex (Z), the cyclometallation process is even more slow and corresponds directly to the reorganization of the ligand to its cyclopalladation-active (E) conformation.  相似文献   

17.
In separation processes, hydrogen bonding has a very significant effect on the efficiency of isolation of acetic acid (HOAc) from HOAc/H2O mixtures. This intermolecular interaction on aggregates composed of a single HOAc molecule and varying numbers of H2O molecules has been examined by using ab initio molecular dynamics simulations (AIMD) and quantum chemical calculations (QCC). Thermodynamic data in aqueous solution were obtained through the self-consistent reaction field calculations and the polarizable continuum model. The aggregation free energy of the aggregates in gas phase as well as in aqueous system shows that the 6-membered ring is the most favorable structure in both states. The relative stability of the ring structures inferred from the thermodynamic properties of the QCC is consistent with the ring distributions of the AIMD simulation. The study shows that in dilute aqueous solution of HOAc the more favorable molecular interaction is the hydrogen bonding between HOAc and H2O molecules, resulting in the separation of acetic acid from the HOAc/H2O mixtures with more difficulty than usual.  相似文献   

18.
6-Acetoxy-4-alkenyl arenes undergo regio- and stereoselective intramolecular Friedel-Crafts reaction affording benzocycles in moderate to excellent yields in TFA/HOAc (3:1). It was observed that introduction of alkyls or phenyl group to the allylic acetate moiety facilitates the cyclization reaction. The optically active tricyclic (4bR,8aS)-4b,7,8,8a,9,10-hexahydrophenanthrene skeleton could also be easily obtained in excellent yields.  相似文献   

19.
The potential energy surface of 1-H-pyrrolo[3,2-b]pyridine-3-yl acetic acid has been investigated via RIIF/6-31G* calculations. The stationary points and reaction paths for syn orientation of the COOH group were determined and are compared with those of the derivatives of 3-indole acetic acid, which act as plant growth hormones. 1-H-pyrrolo[3,2-b]pyridine-3-yl acetic acid forms a kinetically stable conformer with a strong intramolecular hydrogen bond, in which the COOH group is in anti orientation. The influence of this hydrogen bond on bond lengths and vibration frequencies is described.  相似文献   

20.
Conventional kinetic analysis of the reactions of nitromethane (NM), nitroethane (NE) and 2-nitropropane (2-NP) with hydroxide ion in water revealed that the reactions are complex and involve kinetically significant intermediates. Kinetic experiments at the isosbestic points where changes in reactant and product absorbance cancel indicate the evolution and decay of absorbance characteristic of the formation of reactive intermediates. The deviations from 1st-order kinetics were observed to increase with increasing extent of reaction and in the reactant order: NM < NE < 2-NP. The apparent deuterium kinetic isotope effects for proton/deuteron transfer approach unity near zero time and increased with time toward plateau values as the reaction kinetics reach steady state. It is proposed that the initially formed preassociation complexes are transformed to more intimate reactant complexes which can give products by two possible pathways.  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