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 hydrolysis of iminohydantoins generates the same tetrahedral intermediate as that obtained in the cyclization of hydantoic acid amides to hydantoins. The ratio of the products of imine hydrolysis under kinetic control is determined by the relative height of the barriers of the breakdown of to amide or to hydantoin. Thus the partitioning of products unequivocally proves which is the rate determining step in the cyclization reaction-formation or breakdown of . UV and 1H NMR monitoring of the acid catalyzed hydrolysis of four 5-substituted 4-imino-1-methyl-3-(4-nitrophenyl)imidazolidin-2-ones found hydantoins as the only products. The kinetics of hydrolysis of imines were measured in 0.001-1 M HCl. Contrary to the remaining imines, 1,5-dimethyl-4-imino-3-(4-nitrophenyl)imidazolidin-2-one is readily oxidized as stock solution in THF containing peroxides to 1,5-dimethyl-5-hydroxy-4-imino-3-(4-nitrophenyl)imidazolidin-2-one . In all cases, hydrolysis was found to be zero order with respect to [H+]. As imines are fully protonated under the acidity studied, this is evidence of a transition state of a single positive charge. Comparison of imine hydrolysis rates with previous data on rates of cyclization of the corresponding amides of hydantoic acids allowed conditions (acid concentration, substitution pattern-gem-dimethyl effect) to be found that guaranteed kinetic control of the products obtained. Thus it was unequivocally proven that formation of the tetrahedral intermediate is rate determining in the cyclization of hydantoic acid amides. The small steric effects upon methyl substitution at 5-C and a solvent kinetic isotope effect kH/kD of 1.72 favour a mechanism for imine hydrolysis whereby the rate is limited by water attack on the protonated imine concerted with proton transfer from attacking water to a second water molecule.  相似文献   

2.
The reaction between CpFe(dppe)H and a number of different proton donors (2-fluoroethanol, MFE; 2,2,2-trifluoroethanol, TFE; hexafluoro-2-propanol, HFIP; perfluoro-tert-butyl alcohol, PFTB; and trifluoroacetic acid, TFA) has been investigated spectroscopically by variable-temperature infrared, UV-visible, and NMR spectroscopy, and has been measured kinetically by the stopped-flow technique with UV-visible detection. The low-temperature IR study shows the establishment of hydrogen-bonding interactions which involve the hydride ligand as the proton accepting site. This investigation quantifies the thermodynamics of the hydrogen-bonding interaction and the basicity factor (E(j)) of the hydride complex. All techniques agree in indicating an equilibration process, after the immediate hydrogen-bond formation, between the hydride complex and an intermediate dihydrogen complex, [CpFe(dppe)(H(2))](+). The equilibrium is shifted toward the dihydrogen complex to a greater extent for the stronger alcohols and for higher alcohol/Fe ratios. The observed equilibration rate constant is linearly dependent on the alcohol concentration, in agreement with the involvement of two alcohol molecules and the formation of a homoconjugate pair. The rate constant increases with the acidity of the proton donor (TFE < HFIP < PFTB < TFA). The rate of the subsequent irreversible isomerization leading to the classical dihydride complex, [CpFe(dppe)H(2)](+), is first order, and the rate constant does not depend on the proton donor nature. The reaction continues, if conducted in CH(2)Cl(2), with a third, slower step leading to the paramagnetic [CpFe(dppe)Cl](+) product. The kinetic data are in accord with an isomerization mechanism consisting of an intramolecular reorganization, leading in one step from the dihydrogen complex to the classical dihydride species, and disagree with the occurrence of a proton-transfer process at the metal site.  相似文献   

3.
The acid-basic interaction between tetra(4-nitro-5-tert-butyl)phthalocyanine and pyridine, 2-methylpyridine, morpholine, piperidine, n-butylamine, diethylamine, and triethylamine in benzene is studied. It is found that the intermolecular transfer of protons of NH groups from tetra(4-nitro-5-tert-butyl)phthalocyanine to morpholine and diethylamine is characterized by unusually low values of the reaction constant rates. The effect of the structure of tetra(4-nitro-5-tert-butyl)phthalocyanine and tetra(3-nitro-5-tert-butyl)phthalocyanine, and of the nature of the base on the kinetic parameters of acid-base interaction is demonstrated. A structure is proposed for complexes with the transfer of displaced phthalocyanines’ protons. It is found that they undergo decomposition over time.  相似文献   

4.
The cyclization of N-nitrosoaminonitriles was studied by a kinetic method with spectrophotometric recording. It is shown that the reaction rate increases as the electron-donor properties of the substituent increase and as the size of the substituent increases. The rate-determining step is detachment of a proton from the C4 atom in the 4H-1,2,3-oxodiazole-5-imine dication.This study was carried out with a Shimazdu MPS-50L septrophotometer.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 1, pp. 39–46, January, 1977.  相似文献   

5.
《Chemical physics letters》1985,116(4):295-301
The value of the proton diffusion coefficient DH+ in ice was extracted from the diffusion-controlled rate kD of the proton recombination reaction RO + H3O+kD ROH + H2O in polycrystalline doped ice. At −10°C, DH+ was estimated to lie between 3.5×10−6 and 1.3×10−5 cm2 s−1, well below the corresponding value of (4.1 ± 0.1)×10−5 cm2 s−1 found in supercooled water.  相似文献   

6.
Acid-base interactions of hexa(m-trifluoromethylphenyl)benzotetraazaporphine with pyridine, 2-methylpyridine, morpholine, benzylamine, piperidine, n-butylamine, diethylamine, triethylamine, and tert-butylamine in benzene are studied. It is found that intermolecular proton transfer from the NH groups of hexa(m-trifluoromethylphenyl)benzotetraazaporphine to n-butylamine and piperidine is characterized by unusually low values of rate constants. The effect of the structure of hexa(m-trifluoromethylphenyl)benzotetraazaporphine and octa(m-trifluoromethylphenyl)tetraazaporphine, and the nature of the base on the kinetic parameters of acid-base interactions, is demonstrated. A structure for proton-transfer complexes of substituted tetraazaporphines is proposed. It is found that they decompose over time.  相似文献   

7.
Experiments have been performed to elucidate the mechanism of proton transfer in ternary clusters containing the matrix-assisted laser-desorption ionization (MALDI) matrix sinapinic acid, nonchromophoric analytes (proline, methionine, and prolylmethionine), and argon. To investigate the mechanism of intracluster proton transfer, ionizing laser power studies were performed at 266 and 355 nm. Baseline studies show that two photons are required at both wavelengths for the formation of sinapinic acid radical cations from sinapinic acid/argon clusters. Studies of the ternary sinapinic acid/biomolecule/argon clusters show that, in all cases, the photon dependence for protonation of the biomolecule is the same as that for formation of the sinapinic acid radical cation. Furthermore, the slopes of the power plots are generally between 1.5 and 2.0, consistent with a two photon ionization process. No evidence of negative ion formation is detected in the negative ion mass spectra. The combined results are consistent with a mechanism of biomolecular intracluster protonation via proton transfer from the photoionized sinapinic acid radical cation. Wavelength dependent trends in matrix and analyte fragment ion formation in conventional MALDI mass spectra and the cluster proton transfer mass spectra were noted. The possible contribution of cluster proton transfer to the analyte protonation mechanism in conventional MALDI is discussed.  相似文献   

8.
The mechanism of transition-metal tetrahydroborate dimerization was established for the first time on the example of (Ph(3)P)(2)Cu(η(2)-BH(4)) interaction with different proton donors [MeOH, CH(2)FCH(2)OH, CF(3)CH(2)OH, (CF(3))(2)CHOH, (CF(3))(3)CHOH, p-NO(2)C(6)H(4)OH, p-NO(2)C(6)H(4)N═NC(6)H(4)OH, p-NO(2)C(6)H(4)NH(2)] using the combination of experimental (IR, 190-300 K) and quantum-chemical (DFT/M06) methods. The formation of dihydrogen-bonded complexes as the first reaction step was established experimentally. Their structural, electronic, energetic, and spectroscopic features were thoroughly analyzed by means of quantum-chemical calculations. Bifurcate complexes involving both bridging and terminal hydride hydrogen atoms become thermodynamically preferred for strong proton donors. Their formation was found to be a prerequisite for the subsequent proton transfer and dimerization to occur. Reaction kinetics was studied at variable temperature, showing that proton transfer is the rate-determining step. This result is in agreement with the computed potential energy profile of (Ph(3)P)(2)Cu(η(2)-BH(4)) dimerization, yielding [{(Ph(3)P)(2)Cu}(2)(μ,η(4)-BH(4))](+).  相似文献   

9.
The kinetics and mechanism of the reactions of the dimeric and monomeric methyloxo(dithiolato)rhenium(V) complexes [(o-SC6H4CH2S)Re(O)CH3]2 and [(o-SC6H4CH2S)PyRe(O)CH3] (Py = pyridine) with XO, sulfoxides, and pyridine N-oxides are studied. In these reactions, an oxygen atom from XO is transferred to rhenium, from which it later removed. A reaction scheme is proposed to interpret the kinetic data. This scheme features the formation of a monomeric (sulfoxide)- or (pyridine N-oxide)(dithiolato)methyloxorhenium(V) complex followed by its bimolecular oxidation in a rate-controlling step. Several sulfoxides (methyl, methyl phenyl, and substituted diphenyl) all react at similar rates. Activation parameters are determined for dimethyl sulfoxide and di-4-tolyl sulfoxide from temperature-dependent studies. The reactions with pyridine N-oxides show autocatalysis in which the catalyst is confirmed to be pyridine formed in the reactions.  相似文献   

10.
The pressure dependence of the excited-state proton dissociation rate constant of four photoacids, 2-naphthol-6,8-disulfonate (2N68DS), 10-hydroxycamptothecin (10-CPT), 5-cyano-2-naphthol (5CN2), and 5,8-dicyano-2-naphthol (DCN2), are studied in methanol. The results are compared with the results of the pressure dependence study we recently conducted for several photoacids in water, ethanol, and propanol. The pressure dependence is explained using an approximate stepwise two-coordinate proton transfer model. The increase in rate, as a function of pressure, manifests a strong dependence of proton tunneling on the distance which decreases with an increase of pressure between the two oxygen atoms involved in the process. The decrease in the proton transfer rate with increasing pressure reflects the dependence of the reaction on the solvent relaxation rate. We found that, for the relatively weak photoacids 2N68DS, 10-CPT, and 5CN2, the proton transfer rate constant increases by a factor of about 5-8 at a pressure of about 1.5 GPa. For a strong photoacid like DCN2, the rate increase was only by a factor of 2.  相似文献   

11.
Silver(I) catalyzed oxidation of aspartic acid by cerium(IV) was studied in acid perchlorate medium. The stoichiometry of the reaction is represented by the eq. (i) Dimeric cerium(IV) species has been indicated and employed in calculations of monomeric cerium(IV) species concentrations. The reaction is second-order and uncatalyzed reaction also simultaneously occurs along with the silver(I) catalyzed reaction conforming to the rate law (ii) where k is an observed second-order rate constant. A probable reaction mechanism is suggested. © 1995 John Wiley & Sons, Inc.  相似文献   

12.
The sulfur-bridged dimeric dithiolato rhenium(V) chelate [CH3(O)Re(eta 2,mu-o-SCH2C6H4S)]2 (D), derived from 2-mercaptothiophenol, was monomerized to give [CH3(O)Re(eta 2-o-SCH2C6H4S)]L (M-L) in benzene upon reaction with various neutral and anionic monodentate ligands (L) such as pyridine and its substituted derivatives, triarylphosphines, dimethyl sulfoxide, 4-picoline-N-oxide, and halide ions. The kinetic observations can readily be interpreted for all ligands by a unified mechanism in which the initial fast formation of a 1:1 (DL) and 1:2 (DL2) adduct is followed by the slow monomerization of each species so formed. The use of different ligands gave insight into different steps of the same multistep mechanism. The kinetics of ligand exchange between free L and the monomeric complexes was also studied; an associative pathway has been proposed to interpret the results. The crystal structures of two new monomeric ML complexes (with L = 4-acetylpyridine and 1,3-diethylthiourea) are reported.  相似文献   

13.
Zhou L  Wang J  Zhang Y  Yao Y  Shen Q 《Inorganic chemistry》2007,46(14):5763-5772
The synthesis and structures of a series of lanthanide(II) and lanthanide(III) complexes supported by the amido ligand N(SiMe3)Ar were described. Several lanthanide(III) amide chlorides were synthesized by a metathesis reaction of LnCl3 with lithium amide, including {[(C6H5)(Me3Si)N]2YbCl(THF)}2.PhCH3 (1), [(C6H3-iPr2-2,6)(SiMe3)N]2YbCl(mu-Cl)Li(THF)3.PhCH3 (4), [(C6H3-iPr2-2,6)(SiMe3)N]YbCl2(THF)3 (6), and [(C6H3-iPr2-2,6)(SiMe3)N]2SmCl3Li2(THF)4 (7). The reduction reaction of 1 with Na-K alloy afforded bisamide ytterbium(II) complex [(C6H5)(Me3Si)N]2Yb(DME)2 (2). The same reaction for Sm gave an insoluble black powder. An analogous samarium(II) complex [(C6H5)(Me3Si)N]2Sm(DME)2 (3) was prepared by the metathesis reaction of SmI2 with NaN(C6H5)(SiMe3). The reduction reaction of ytterbium chloride 4 with Na-K alloy afforded monoamide chloride {[(C6H3-iPr2-2,6)(SiMe3)N]Yb(mu-Cl)(THF)2}2 (5), which is the first example of ytterbium(II) amide chloride, formed via the cleavage of the Yb-N bond. The same reduction reaction of 7 gave a normal bisamide complex [(C6H3-iPr2-2,6)(SiMe3)N]2Sm(THF)2 (8) via Sm-Cl bond cleavage. This is the first example for the steric effect on the outcome of the reduction reaction in lanthanide(II) chemistry. 5 can also be synthesized by the Na/K alloy reduction reaction of 6. All of the complexes were fully characterized including X-ray diffraction for 1-7.  相似文献   

14.
In the proton transfer reactions between [Fe 4Y 4Cl 4] (2-) (Y = S or Se) and [pyrH] (+) (pyr = pyrrolidine) in the presence of a variety of nucleophiles (L = I (-), Br (-), PhS (-), EtS (-) or ButNC), initial binding of the nucleophile can occur to generate [Fe 4Y 4Cl 4(L)] ( n- ). The subsequent rate of proton transfer depends markedly on the nature of L. Stopped-flow kinetic studies show that proton transfer from [pyrH] (+) to [Fe 4Y 4Cl 4] (2-) { (S) k 4 = (2.1 +/- 0.5) x 10 (4) dm (3) mol (-1) s (-1); (Se) k 4 = (8.0 +/- 0.5) x 10 (3) dm (3) mol (-1) s (-1)} is increased by prior binding of L = PhS (-) or Bu ( t )NC to form [Fe 4Y 4Cl 4(L)] (n-) ( (S) k 7 (L) approximately 1 x 10 (6) dm (3) mol (-1) s (-1)), but prior binding of L = I (-), Br (-), or EtS (-) to the clusters inhibits the rate of proton transfer {e.g. (S) k 7 (I) = (6.0 +/- 0.8) x 10 (2) dm (3) mol (-1) s (-1); (Se) k 7 (I) = (4.5 +/- 0.5) x 10 (2) dm (3) mol (-1) s (-1)}. This behavior is correlated with the bonding characteristics of L and the effect this has on bond length reorganization within the cluster upon proton transfer.  相似文献   

15.
Intracluster proton transfer from the matrix-assisted laser desorption/ionization matrix 2,5-dihydroxybenzoic acid (DHB) to the peptide valyl-prolyl-leucine has been investigated as a function of excitation laser wavelength and power. Ionization laser power studies at 308 nm indicate that cluster ionization occurs with a two-photon dependence, whereas matrix-to-analyte proton transfer and cluster dissociation requires an additional photon. At 266 nm, two-photon absorption leads to both cluster ionization and cluster dissociation/proton transfer. A consideration of these results clearly indicates that analyte protonation occurs following ionization of the cluster to produce a radical cation matrix/analyte cluster. Mass spectral features also indicate that mixed DHB/peptide cluster ionization can occur via two-photon ionization at wavelengths as long as 355 nm. These results suggest a reduction in the ionization potential of larger mixed DHB/peptide clusters of greater than 1 eV. The reduced ionization potential seen in these clusters suggests that radical cation initiated proton transfer remains a viable mechanism for analyte protonation in matrix-assisted laser desorption/ionization at these longer wavelengths.  相似文献   

16.
The kinetics of oxidation of Aspartic acid by bismuth(V) studied in HClO4-HF mixture iodometrically exhibit complex dependence with respect to aspartic acid (AA). The rate law (i) accounts for all the experimental observations. where [Bi(V)] and [AA] are the gross analytical concentration of all fluorobismuth(V) species and the equilibrium concentration of aspartic acid, respectively. The oxidation product of the amino acid was identified to be an aldehyde. HF and F? do not affect the rate of the reaction. © John Wiley & Sons, Inc.  相似文献   

17.
The kinetics of the interactions of the Keggin type heteropolyanion [CoW12O40]5?, with all the molecular and anionic forms of citric acid have been investigated spectrophotometrically in the pH range 0.60–5.45 at 60°C. With the exception of the molecular form of the acid (H3L), all the other species (H2L?, HL2?, and L3?) undergo oxidation through an alkali metal ion catalyzed path in addition to the uncatalyzed one. The catalytic power increases with increasing size of the alkali metal ion. The contributions of each path have been evaluated from the kinetic data with the help of the measured dissociation constant values of citric acid. A plausible mechanism involving the formation of a bridged outer-sphere complex of the reacting species with the alkali cation, has been suggested. The observed limiting dependence of kobs on [citrate] at high concentrations owing to an equilibrium between the reactants preceeding the electron transfer step supports the proposed mechanism. The pH-rate profile gives an excellent fit with the evaluated kinetic parameters.  相似文献   

18.
Summary The reaction between sulphurous acid and hexachloroiridate(IV) appears to take place through the formation of an intermediate complex followed by decomposition to give oxidation products. The rate is retarded as the hydrogen ion concentration increases. Thermodynamic parameters associated with the equilibrium step as well as with the slowest step have been calculated. The probable mechanism of the reaction is discussed.  相似文献   

19.
The intramolecular mechanism we earlier proposed [Alagona, G.; Desmeules, P.; Ghio, C.; Kollman, P. A. J Am Chem Soc 1984, 106, 3623] for the second proton transfer of the reaction catalyzed by triosephosphate isomerase (TIM) is examined ab initio at the HF and MP2/6-31+G** levels in vacuo for two conformers of the enediolate phosphate (ENEP), with the ethereal oxygen of the phosphate group either syn (X), as in the crystal structure, or anti (Y) with respect to the enediolate carbonyl O. The barrier height for the intramolecular proton transfer occurring in enediolate is very sensitive to electron correlation corrections. The MP2 internal energy barrier is much lower than the HF one, while the free energy (FE) barrier is even more favorable, indicating that the enzyme presence is not requested to speed up that step. An investigation of the dynamical aspects of the mechanism, along the pathway from ENEP A (with H on O(1)) to TS and from TS to ENEP B (with H on O(2)), was, however, carried out in the presence of the enzyme field while using a neutral His-95 with its proton on Ndelta. To perform the FE simulations, it was necessary to parametrize in the AMBER force-field the ENEP A, TS and B species, whose partial charges have been determined with the RESP procedure, with the X and Y arrangements of the phosphate head. Actually, the FE/QM approach produced a low barrier and a substantial balance between A and B, especially at the MP2 level. The trajectories, analyzed paying a particular attention to the positions assumed by His-95 and by the other active site residues, put forward somewhat different H-bond patterns around the X or Y enediolate phosphate.  相似文献   

20.
The sub-millisecond protonation dynamics of the chromophore in S65T mutant form of the green fluorescent protein (GFP) was tracked after a rapid pH jump following laser-induced proton release from the caged photolabile compoundo-nitrobenzaldehyde. Following a jump in pH from 8 to 5 (which is achieved within 2 μs), the fluorescence of S65T GFP decreased as a single exponential with a time constant of ∼90 μs. This decay is interpreted as the conversion of the deprotonated fluorescent GFP chromophore to a protonated non-fluorescent species. The protonation kinetics showed dependence on the bulk viscosity of the solvent, and therefore implicates bulk solvent-controlled protein dynamics in the protonation process. The protonation is proposed to be a sequential process involving two steps: (a) proton transfer from solvent to the chromophore, and (b) internal structural rearrangements to stabilize a protonated chromophore. The possible implications of these observations to protein dynamics in general is discussed  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