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 kinetics of methanolysis of six O-ethyl O-aryl methylphosphonates (6a-f) promoted by methoxide, La3+ and 1,5,9-triazacyclododecane complex of Zn2+(-OCH3) (5:Zn2+(-OCH3)) were studied as simulants for chemical warfare (CW) agents, and analyzed through the use of Br?nsted plots. The beta(lg) values are, respectively, -0.76, -1.26 and -1.06, pointing to significant weakening of the P-OAr bond in the transition state. For the metal-catalyzed reactions the data are consistent with a concerted process where the P-OAr bond rupture has progressed to the extent of 84% in the La3+ reaction and ca. 70% in the Zn2+ catalyzed reaction. The catalysis afforded by the metal ions is remarkable, being about 10(6)-fold and 10(8)-fold for poor and good leaving groups, respectively, relative to the background reactions at pH 9.1. Solvent deuterium kinetic isotope studies for two of the substrates promoted by 5:Zn2+(-OCH3) give kH/kD = 1.0 +/- 0.1, consistent with a nucleophilic mechanism. A unified mechanism for the metal-catalyzed reactions is presented which involves pre-equilibrium coordination of the substrate to the metal ion followed by intramolecular delivery of a coordinated methoxide.  相似文献   

2.
The methanolyses of a series of O,O-diethyl O-aryl phosphates (2,5) and O,O-diethyl S-aryl phosphorothioates (6) promoted by methoxide and two metal ion systems, (La3+)2(-OCH3)2 and 4:Zn2+:-OCH3 (4 = 1,5,9-triazacyclododecane) has been studied in methanol at 25 degrees C. Br?nsted plots of the logk2 values vs. pKa for the phenol leaving groups give beta(lg) values of -0.70, -1.43 and -1.12 for the methanolysis of the phosphates and -0.63, -0.87 and -0.74 for the methanolysis of the phosphorothioates promoted by the methoxide, La3+ and Zn2+ systems respectively. The kinetic data for the metal-catalyzed reactions are analyzed in terms of a common mechanism where there is extensive cleavage of the P-XAr bond in the rate-limiting transition state. The relevance of these findings to the mechanism of action of the phosphotriesterase enzyme is discussed.  相似文献   

3.
Ion-molecule reactions between the O=P(OCH(3))(2) (+) phosphonium ions and five aliphatic esters (methyl acetate, methyl propionate, methyl 2-methylpropionate, methyl butyrate and ethyl acetate) were performed in a quadrupole ion trap mass spectrometer. The O=P(OCH(3))(2) (+) phosphonium ions, formed by electron ionization from neutral trimethyl phosphite, were found to react with aliphatic esters to give an adduct ion [RR'CHCOOR", O=P(OCH(3))(2)](+), which loses spontaneously a molecule of ketene CH(2)=CO or substituted ketenes RR'C=CO. Isotope-labeled methyl acetate was used to elucidate fragmentation mechanisms. The potential energy surface obtained from B3LYP/6-31G(d,p) calculations for the reaction between O=P(OCH(3))(2) (+) and methyl acetate is described.  相似文献   

4.
The di-Zn(II) complex of 1,3-bis[ N1, N1'-(1,5,9-triazacyclododecyl)]propane with an associated methoxide ( 3:Zn(II) 2: (-)OCH 3) was prepared and its catalysis of the methanolysis of a series of fourteen methyl aryl phosphate diesters ( 6) was studied at s (s)pH 9.8 in methanol at 25.0 +/- 0.1 degrees C. Plots of k obs vs [ 3:Zn(II) 2: (-)OCH 3] free for all members of 6 show saturation behavior from which K(M) and kcat (max) were determined. The second order rate constants for the catalyzed reactions (kcat (max)/K(M)) for each substrate are larger than the corresponding methoxide catalyzed reaction (k 2 (-OMe)) by 1.4 x 10(8) to 3 x 10 (9)-fold. The values of k cat (max) for all members of 6 are between 4 x 10(11) and 3 x 10(13) times larger than the solution reaction at s (s)pH 9.8, with the largest accelerations being given for substrates where the departing aryloxy unit contains ortho-NO 2 or C(O)OCH 3 groups. Based on the linear Br?nsted plots of k cat (max) vs s (s)pKa of the phenol, beta lg values of -0.57 and -0.34 are determined respectively for the catalyzed methanolysis of "regular" substrates that do not contain the ortho-NO 2 or C(O)OCH 3 groups, and those substrates that do. The data are consistent with a two step mechanism for the catalyzed reaction with rate limiting formation of a catalyst-coordinated phosphorane intermediate, followed by fast loss of the aryloxy leaving group. A detailed energetics calculation indicates that the catalyst binds the transition state comprising [CH 3O (-): 6], giving a hypothetical [ 3:Zn(II) 2:CH 3O (-): 6] complex, by -21.4 to -24.5 kcal/mol, with the strongest binding being for those substrates having the ortho-NO 2 or C(O)OCH 3 groups.  相似文献   

5.
The methanolyses of several organophosphate/phosphonate/phosphorothioate esters (O,O-diethyl O-(4-nitrophenyl) phosphate, paraoxon, ; O,O-diethyl S-(3,5-dichlorophenyl) phosphorothioate, ; O-ethyl O-(2-nitro-4-chlorophenyl) methylphosphonate, ; O,O-dimethyl O-(3-methyl-4-nitrophenyl) phosphorothioate, fenitrothion, ; O-ethyl S-(3,5-dichlorophenyl) methylphosphonothioate ) and a carboxylate ester (p-nitrophenyl acetate, ) catalyzed by methoxide and the Zn(2+)((-)OCH(3)) complex of 1,5,9-triazacyclododecane ( : Zn(2+)((-)OCH(3))) were studied in methanol and d(1)-methanol at 25 degrees C. In the case of the methoxide reactions inverse skie's were observed for the series with values ranging from 2 to 1.1, except for where the k(D)/k(H) = 0.90 +/- 0.02. The inverse k(D)/k(H) values are consistent with a direct nucleophilic methoxide attack involving desolvation of the nucleophile with varying extents of resolvation of the TS. With the : Zn(2+)((-)OCH(3)) complex all the skie values are k(D)/k(H) = 1.0 +/- 0.1 except for where the value is 0.79 +/- 0.06. Arguments are presented that the fractionation factors associated with complex : Zn(2+)((-)OCH(3)) are indistinguishable from unity. The skie's for all the complex-catalyzed methanolyses are interpreted as being consistent with an intramolecular nucleophilic attack of the Zn(2+)-coordinated methoxide within a pre-equilibrium metal : substrate complex.  相似文献   

6.
The La3+-catalyzed methanolysis of two phosphorothioate derivatives, O,O-diethyl S-(p-nitrophenyl) phosphorothioate (4a) and O,O-diethyl S-phenyl phosphorothioate (4b) were studied as a function of [La3+] and pH in methanol solvent. In both cases the kinetics of catalyzed methanolysis maximize at pH 9.1 and a detailed analysis indicates that the dominant species responsible for catalysis are dimers formulated as La3+(2)(-OCH3)2 and La3+(2)(-OCH3)4. The catalysis is compared with that seen for the corresponding phosphate esters, namely paraoxon (3a) and O,O-diethyl phenyl phosphate (3b) for which La3+ catalysis is slightly better and markedly worse than for 4a and 4b respectively. Overall, at s(s)pH 9.1, a 2 mmol dm-3 solution of La(OTf)3 with equimolar NaOCH3 provides accelerations of 2.2x10(8)-fold, 9.7x10(6)-fold and 9.3x10(6)-fold for methanolysis of 3a, 4a and 4b, relative to the background reaction of methoxide reacting with the three substrates. In each case, the P-containing product of the reactions is exclusively diethyl methyl phosphate. Turnover experiments with 6-fold and 100-fold excesses of 4a and 4b respectively, methanolyzed in the presence of approximately 10 mmol dm-3 La3+ and equimolar NaOCH3, indicate that the reactions are essentially complete within 103 s and 70 min respectively. The latter turnover experiment with 4b corresponded to 100 turnovers in 70 min and an overall reaction t1/2 of 8 min. A common mechanism of reaction is postulated for each of the substrates which involves Lewis acid coordination of one of the La3+ to the P=O unit, followed by nucleophilic attack by the second La3+-(-)OCH3.  相似文献   

7.
Ion-molecule reactions between the O=P(OCH3)2 + phosphonium ions and eight alpha,beta-unsaturated esters (methyl acrylate, ethyl acrylate, methyl crotonate, ethyl crotonate, methyl 3,3-dimethylacrylate, ethyl 3,3-dimethylacrylate, methyl methacrylate and ethyl methacrylate) were performed in a quadrupole ion trap mass spectrometer. The O=P(OCH3)2 + phosphonium ions, formed by electron ionization from neutral trimethyl phosphite, were found to react with alpha,beta-unsaturated esters to give an adduct [RR'C=CR'COOR', O=P(OCH3)2]+, which lose spontaneously a molecule of trimethyl phosphate (R'=CH3) or dimethyl ethyl phosphate (R'=C2H5). An ion corresponding to a protonated trialkyl phosphate is also observed when substituent R'=H. To confirm the experimental results, and to elucidate the mechanism for the formation of the ionic species, a theoretical study using the density functional theory (DFT) approach was carried out. The potential energy surface obtained from B3LYP/6-31G(d,p) calculations for the reaction between O=P(OCH3)2 + and methyl acrylate is described.  相似文献   

8.
This paper reports on a spectrophotometric kinetic study of the effects of the alkali metal ions Li+ and K+ on the ethanolysis of the aryl methyl phenyl phosphinate esters 3a-f in anhydrous ethanol at 25 degrees C. Rate data obtained in the absence and presence of complexing agents afford the second-order rate constants for the reaction of free ethoxide (k(EtO-)) and metal ion-ethoxide ion pairs (k(MOEt)). The sequence k(EtO-) < k(MOEt) is established for all the substrates, contrary to the generally observed reactivity order in nucleophilic substitution processes. The quantities deltaG(ip), deltaG(ts) and DeltaG(cat), which quantify the observed alkali metal ion effect in terms of transition state stabilization through chelation of the metal ion, give the order deltaG(ts) > deltaG(ip) for Li+ and K+. Hammett plots show significantly better correlation of rates with sigma and sigma(o) substituent constants than with sigma-, yielding moderately large rho(rho(o)) values that are consistent with a stepwise mechanism in which formation of a pentacoordinate (phosphorane) intermediate is the rate-limiting step. The range of the values of the selectivity parameter, rho(n) (= rho]/rho(eq)), 1.3-1.6, obtained for the uncatalyzed and alkali metal ion catalyzed reactions indicates that there is no significant perturbation of the transition state (TS) structure upon chelation of the metal ions. This finding is relevant to the mechanism of enzymatic phosphoryl transfer involving metal ion co-factors. The present results enable one to compare structural effects for nucleophilic reactions of several series of organophosphorus substrates. It is shown that the order of reactivity of the substrates: 4-nitrophenyl dimethyl phosphinate (2) > 3a > 4-nitrophenyl diphenyl phosphinate (1) is determined mainly by the steric effects of the alkyl/aryl substituents around the central P atom in the TS of the reaction.  相似文献   

9.
A methoxide-bridged dinuclear Zn(II) complex of 1,3-[N,N'-bis(1,5,9-triazacyclododecane)]propane (1-Zn(II)2:(-OCH3)) was prepared, and its catalysis of the cyclization of a series of 2-hydroxypropyl aryl phosphates (4a-g) was investigated in methanol at pH 9.8, T = 25degreesC by stopped-flow spectrophotometry. An X-ray diffraction structure of the hydroxide analogue of 1-Zn(II)2:(-OCH3), namely 1-Zn(II)2:(-OH), reveals that each of the Zn(II) ions is coordinated by the three N's of the triazacyclododecane units and a bridging hydroxide. The cyclizations of substrates 4a-g reveal a progressive change in the observed kinetics from Michaelis-Menten saturation kinetics for the poorer substrates (4-OCH3 (4g); 4-H (4f); 3-OCH3 (4e); 4-Cl (4d); 3-NO2, (4c)) to second-order kinetics (linear in 1-Zn(II)2:(-OCH3)) for the better substrates (4-NO2,3-CH3 (4b); 4-NO2, (4a)). The data are analyzed in terms of a multistep process whereby a first formed complex rearranges to a reactive complex with a doubly activated phosphate coordinated to both metal ions. The kinetic behavior of the series is analyzed in terms of change in rate-limiting step for the catalyzed reaction whereby the rate-limiting step for the poorer substrates (4g-c) is the chemical step of cyclization of the substrate, while for the better substrates (4b,a) the rate-limiting step is binding. The catalysis of the cyclization of these substrates is extremely efficient. The kcat/KM values for the catalyzed reactions range from 2.75 x 10(5) to 2.3 x 10(4) M-1 s-1, providing an acceleration of 1 x 10(8) to 4 x 10(9) relative to the methoxide reaction (k2OCH3, which ranges from 2.6 x 10(-3) to 5.9 x 10(-6) M-1 s-1 for 4a-g). At a pH of 9.8 where the catalyst is maximally active, the acceleration for the substrates ranges from (1 - 4) x 10(12) relative to the background reaction at the same pH. Detailed energetics calculations show that the transition state for the catalyzed reaction comprising 1-Zn(II)2, methoxide, and 4 is stabilized by about -21 to -23 kcal/mol relative to the transition state for the methoxide reaction. The pronounced catalytic activity is attributed to a synergism between a positively charged catalyst that has high affinity for the substrate and for the transition state for cyclization, and a medium effect involving a reduced polarity/dielectric constant that complements a reaction where an oppositely charged reactant and catalyst experience charge dispersal in the transition state.  相似文献   

10.
The methanolyses of two neutral phosphorus triesters, paraoxon (1) and fenitrothion (3), were investigated as a function of added Zn(OTf)(2) or Zn(ClO(4))(2) in methanol at 25 degrees C either alone or in the presence of equimolar concentrations of the ligands phenanthroline (4), 2,9-dimethylphenanthroline (5), and 1,5,9-triazacyclododecane (6). The catalysis requires the presence of methoxide, and when studied as a function of added NaOCH(3), the rate constants (k(obs)) for methanolysis of Zn(2+) alone or in the presence of equimolar 4 or 5 maximize at different [(-)OCH(3)]/[Zn(2+)](total) ratios of 0.3, 0.5, and 1.0, respectively. Plots of k(obs) vs [Zn(2+)](total) either alone or in the presence of equimolar ligands 4 and 5 at the [(-)OCH(3)]/[Zn(2+)](total) ratios corresponding to the rate maxima are curved and show a nonlinear dependence on [Zn(2+)](total). In the cases of 4 and 5, this is explained as resulting from formation of a nonactive dimer, formulated as a bis-mu-methoxide-bridged form (L:Zn(2+)((-)OCH(3))(2)Zn(2+):L) in equilibrium with an active monomeric form (L:Zn(2+)((-)OCH(3))). In the case of the Zn(2+):6 system, no dimeric forms are present as can be judged by the strict linearity of the plots of k(obs) vs [Zn(2+)](total) in the presence of equimolar 6 and (-)OCH(3). Analysis of the potentiometric titration curves for Zn(2+) alone and in the presence of the ligands allows calculation of the speciation of the various Zn(2+) forms and shows that the binding to ligands 4 and 6 is very strong, while the binding to ligand 5 is weaker. Overall the best catalytic system is provided by equimolar Zn(2+), 5, and (-)OCH(3), which exhibits excellent turnover of the methanolysis of paraoxon when the substrate is in excess. At a concentration of 2 mM in each of these components, which sets the pH of the solution at 9.5, the acceleration of the methanolysis of paraoxon and fenitrothion relative to the methoxide reaction is 1.8 x 10(6)-fold and 13 x 10(6)-fold, respectively. A mechanism for the catalyzed reactions is proposed which involves a dual role for the metal ion as a Lewis acid and source of nucleophilic Zn(2+)-bound (-)OCH(3).  相似文献   

11.
A family of triethanolamine complexes of titanium with varying metal/ligand ratios have been prepared from reactions of titanium tetraisopropoxide with triethanolamine. Three nonhydrolytic products, having essentially all isopropoxide ligands substituted by triethanolamine, were prepared as hygroscopic, glassy solids. Crystals of two hexameric titanatrane partial hydrolysis analogues [Ti3(mu 2-O)((HOCH2CH2)2NCH2CH2O)(OCH2CH2)2(mu 2-OCH2CH2)N)2(OCH2CH2)(mu 2- OCH2CH2)2N)]2 (1), and [Ti3(mu 2-O)(OCH(CH3)2)((OCH2CH2)2(mu 2-OCH2CH2)N)2(OCH2CH2)(mu 2- OCH2CH2)2N)]2 (2) were isolated and structurally characterized. The structures consist of a central core of two oxo-bridged dititanatranes (TEA)TiOTi(TEA) (TEAH3 = triethanolamine) with the nonhydrolytic residue (TEA)Ti(TEAH2) included as an adduct in (1), analogously to (TEA)Ti(OPri) in (2).  相似文献   

12.
The kinetics of the La(3+)-catalyzed methanolysis of N-phenyl-beta-lactam (2) and N-p-nitrophenyl-beta-lactam (3) as well as that of nitrocefin (1) were studied at 25 degrees C under buffered conditions. In the case of 2 and 3, the observed second-order rate constants (k(2)(obs)) for catalysis plateau at pH 7.5-7.8, reaching values of 1 x 10(-)(2) and 35 x 10(-)(2) M(-)(1) s(-)(1) respectively. Potentiometric titrations of solutions of 2 x 10(-)(3) M La(OTf)(3) were analyzed in terms of a dimer model (La(3+)(2)((-)OCH(3))(n)()), where the number of methoxides varies from 1 to 5. The species responsible for catalysis in the pH range investigated contain 1-3 methoxides, the one having the highest catalytic activity being La(3+)(2)((-)OCH(3))(2), which comprises 80% of the total La(3+) forms present at its pH maximum of 8.9. The catalysis afforded by the La(3+) dimers at a neutral pH is impressive relative to the methoxide reactions: at pH 8.4 a 1 mM solution of catalyst (generated from 2 mM La(OTf)(3)) accelerated the methanolysis of 2 by approximately 2 x 10(7)-fold and 3 by approximately 5 x 10(5)-fold. As a function of metal ion concentration, the La(3+)-catalyzed methanolysis of 1 proceeds by pathways involving first one bound metal ion and then a second La(3+) leading to a plateau in the k(obs) vs [La(3+)](total) plots at all pH values. The k(max)(obs) pseudo-first-order rate constants at the plateaus, representing the spontaneous methanolysis of La(3+)(2)(1(-)) forms, has a linear dependence on [(-)OCH(3)] (slope = 0.84 +/- 0.05 if all pH values are used and 1.02 +/- 0.03 if all but the two highest pH values are used). The speciation of bound 1 at a La(3+) concentrations corresponding to that of the onset of the kinetic plateau region was approximated through potentiometric titration of the nonreactive 3,5-dinitrobenzoic acid in the presence of 2 equiv of La(OTf)(3). A total speciation diagram for all bound forms of La(3+)(2)(1(-))((-)OCH(3))(n)(), where n = 0-5, was constructed and used to determine their kinetic contributions to the overall pH vs k(max)(obs) plot under kinetic conditions. Two kinetically equivalent mechanisms were analyzed: methoxide attack on La(3+)(2)(1(-))((-)OCH(3))(n)(), n = 0-2; unimolecular decomposition of the forms La(3+)(2)(1(-))((-)OCH(3))(n)(), n = 1-3.  相似文献   

13.
The scope of the barium salt of p-tert-butylcalix[4]arene-crown-5 as a transacylation catalyst has been defined by evaluating its efficiency in the methanolysis of a series of aryl acetates at 25.0 degrees C in MeCN/MeOH 9:1 (v/v) under slightly basic conditions. In this system a phenolic hydroxyl is the acyl-receiving and -releasing unit in a double-displacement mechanism. The complexed barium ion acts both as a nucleophile carrier and a built-in Lewis acid in providing electrophilic assistance to the ester carbonyl both in the acylation and deacylation step (nucleophilic-electrophilic catalysis). Turnover capability is ensured by the acylated intermediate reacting with the solvent more rapidly than the original ester, but a serious drawback derives from the incursion of back-acylation of the liberated phenol. A gradual shift from rate-determining deacylation (p-nitrophenyl acetate) to rate-determining acylation (phenyl acetate) is observed along the investigated series. It is shown that the scope of the catalyst is restricted to acetate esters whose reactivity lies in the range approximately defined by the phenyl acetate-p-nitrophenyl acetate pair, with a maximum efficiency for p-chlorophenyl acetate. Moreover, the catalyst effectively promotes ester interchange between phenols, showing that its activity is not limited to solvolysis reactions. The very high sensitivity of the rate of acylation of the catalyst to leaving group basicity has been interpreted as due to rate-determining decomposition of the tetrahedral intermediate, which is believed to arise from the presumably low basicity of the metal ion stabilized nucleophile. The turnover frequency was in the range of 3.8 x 10(-4) min(-1) for phenyl acetate to 7.4 x 10(-3) min(-1) for p-nitrophenyl acetate ([ArOAc]0=4.0 mM]). A first attempt to enhance the rate of acylation of the catalyst through intramolecular general acid catalysis is also described.  相似文献   

14.
The Ru3(CO)12-catalyed cross-coupling reaction of esters with organoboron compounds leading to ketones is described. A wide variety of functional groups can be tolerated under the reaction conditions. Aromatic boronates function as a coupling partner to give aryl ketones. Acyl-alkyl coupling to dialkyl ketones is also achieved by the use of 9-alkyl-9-BBN in place of boronates. The Ru3(CO)12-catalyzed decarbonylative reduction of esters with ammonium formate (HCOONH4) leading to hydrocarbons is also described. No expected aldehydes are produced, and controlled experiments indicate that aldehydes are not intermediate for the transformation. A hydrosilane can also be used as a reducing reagent in place of HCOONH4. A wide variety of functional groups are compatible for both reactions. The key step for both catalytic reactions is the directing group-promoted cleavage of an acyl carbon-oxygen bond in esters, leading to the generation of acyl transition metal alkoxo complexes.  相似文献   

15.
We report the results of a selected ion flow tube (SIFT) study of the reactions of H3O+, NO+ and O+2 with some nine carboxylic acids and eight esters. We assume that all the exothermic proton transfer reactions of H3O+ with all the acid and esters molecules occur at the collisional rate, i.e. the rate coefficients, k, are equal to kc; then it is seen that k values for most of the NO+ and O+2 reactions also are equal to or close to kc. The major ionic products of the H3O+ reactions with both the acids and esters are the protonated parent molecules, MH+, but minor channels are also evident, these being the result of H2O elimination from the excited (MH+)1 in some of the acid reactions and an alcohol molecule elimination (CH3OH or C2H5OH) in some of the ester reactions. The NO+ reactions with the acids and esters result in both ion-molecule association producing NO+M in parallel with hydroxide ion (OH) transfer with some of the acids, and parallel methoxide ion (CH3O) and ethoxide ion (C2H5O) transfer as appropriate with some of the esters. The O+2 reactions proceed by dissociative charge transfer with the production of two or more ionic fragments of the parent molecules, the different isomeric forms of both the acid and the ester molecules resulting in different product ions.  相似文献   

16.
Transesterification reactions of acyclic oligoether esters E3-E10 with metal alkoxides were accelerated upon noncovalent complexation of the esters with metal ions. In the reaction of monovalent alkaline metal alkoxides, CH(3)ONa and CH(3)OK, plots of the observed rate constants k(obs) with respect to the chain length of E3-E10 showed selective acceleration of the transesterification. Compared with the shortest E3, which can hardly bind metal ion, 4.3- and 6.6-fold accelerations in the maxima were achieved in the combinations of E5/CH(3)ONa and E6/CH(3)OK, respectively. Supramolecular intermediate complex could be spectrometrically visualized by ESI-FT-ICR-MS in the course of reaction. Kinetic experiments, together with structural analyses by means of NMR, mass spectrometry, and DFT calculations of the intermediate complexes, indicate that a size-fit complex of host substitute with alkali metal ion allows strong electron withdrawing due to the close contact of the carbonyl oxygen to the metal ion, resulting in the selective rate enhancement of the reaction, while in the reaction of E3-E10 with a divalent alkaline earth metal alkoxide, (CH(3)CH(2)O)(2)Ba, the k(obs) values increased stepwise with elongation of the side arm to attain an dramatic large acceleration. In comparison with the k(obs) of E3, 4610-fold acceleration was achieved in the reaction of E10. The double activation of the host substrate and guest counter nucleophile at once brings about this extraordinary rate acceleration. The strong wrapping complexation of long oligoether ester with barium ethoxide allows for the effective electron withdrawal from the ester carbonyl group (host activation) as well as separation of the accompanying guest alkoxide anions (guest activation).  相似文献   

17.
The second-order rate constants for the hydrolysis of nitrophenyl esters catalysed by a number of folded designed polypeptides have been determined, and 1900-fold rate enhancements over those of the 4-methylimidazole-catalysed reactions have been observed. The rate enhancements are much larger than those expected from the pKa depression of the nucleophilic His residues alone. Kinetic solvent isotope effects were observed at pH values lower than the pKa values of the leaving groups and suggests that general-acid catalysis contributes in the pH range where the leaving group is predominantly protonated. In contrast, no isotope effects were observed at pH values above the pKa of the leaving group. A Hammett rho value of 1.4 has been determined for the peptide-catalysed hydrolysis reaction by variation of the substituents of the leaving phenol. The corresponding values for the imidazole-catalysed reaction is 0.8 and for phenol dissociation is 2.2. There is therefore, very approximately, half a negative charge localised on the phenolate oxygen in the transition state in agreement with the conclusion that transition-state hydrogen-bond formation may contribute to the observed catalysis. The elucidation at a molecular level of the principles that control cooperativity in the biocatalysed ester-hydrolysis reaction represents the first step towards a level of understanding of the concept of cooperativity that may eventually allow us to design tailor-made enzymes for chemical reactions not catalysed by nature.  相似文献   

18.
The reaction of imidazole in aqueous solution with toluene-4-sulfonate salts of substituted phenyl N-methylpyridinium-4-carboxylate esters obeys the rate law: k(obs) - k(background) = k2[Im] + k3[Im]2 where [Im] is the imidazole concentration present as free base. The parameters k2 and k3 fit Br?nsted type free energy correlations against the pKa of the leaving phenol with betaLg values of -0.65 and -0.42 respectively. The imidazolysis is insensitive to catalysis by general bases and yet k3 for the 3-cyanophenyl ester possesses a deuterium oxide solvent isotope effect of 4.43 consistent with rate limiting proton transfer. A special catalytic function is proposed for decomposition of the tetrahedral addition intermediate (T+/-) via k3 whereby the catalytic imidazole interacts electrophilically with the leaving phenolate ion and removes a proton from the nitrogen in the rate limiting step with subsequent non-rate limiting ArO-C bond fission. This is consistent with the change in effective charge on the leaving oxygen in the transition structure of k3 which is more positive (-0.42) than that expected (-0.60) for the equilibrium formation of the zwitterion intermediate. The catalytic function at the leaving oxygen is likely to be an electrophilic role of the NH as a hydrogen bond donor. In the k2 step the deuterium oxide solvent isotope effect of 1.51 for the 3-cyanophenyl ester and the betaLg of -0.65 are consistent with rate limiting expulsion of the phenolate ion from the T+/- intermediate. The absence of general base catalysis of imidazolysis rules out the established mechanism for aminolysis of esters where T+/- is stabilised by a standard rate limiting proton transfer. The kinetically equivalent term for k3 where T- reacts with the imidazolium ion as an acid catalyst would require this step to be rate limiting and involve proton transfer not consistent with departure of the good aryl oxide leaving group.  相似文献   

19.
The amination of aryl halides in the presence of inexpensive and air-stable alkali metal hydroxide bases and Pd[P(t-Bu)3]2 as catalyst gave arylamines in high yields. The reactions were conducted with a catalytic amount of cetyltrimethylammonium bromide as phase-transfer agent and either aqueous hydroxide or solid hydroxide in the presence of water. This combination of alkali metal hydroxide base, H2O, and the ammonium salt performed as well as NaO-t-Bu in the amination of p-chlorotoluene with dibutylamine. Hydroxide base was suitable for reactions of a wide range of aryl chlorides and bromides with aliphatic and aromatic amines. Some functional groups that were intolerant of tert-butoxide base, such as esters, enolizable ketones, nitriles, and nitro groups, were tolerated by the combination of hydroxide base, H2O, and cetyltrimethylammonium bromide in toluene solvent.  相似文献   

20.
We study the dissociation of water coordinated to a divalent metal ion center, M2+ = Mg2+, Zn2+ using density functional theory (DFT) and ab initio molecular dynamics (AIMD) simulations. First, the proton affinity of a coordinated OH- group is computed from gas-phase Mg2+(H2O)5(OH-), which yields a relative higher gas-phase acidity for a Zn2+-coordinated as compared to a Mg2+-coordinated water molecule, DeltapKa(gp) = 5.3. We explain this difference on the basis of a gain in stabilization energy of the Zn2+(H2O)5(OH-) system arising from direct orbital interaction between the coordinated OH- and the empty 4s state of the cation. Next, we compute the acidity of coordinated water molecules in solution using free-energy thermodynamic integration with constrained AIMD. This approach yields pKa Mg2+ = 11.2 and pKa Zn2+ = 8.4, which compare favorably to experimental data. Finally, we examine the factors responsible for the apparent decrease in the relative Zn2+-coordinated water acidity in going from the gas-phase (DeltapKa(gp) = 5.3) to the solvated (DeltapKa = 2.8) regime. We propose two simultaneously occurring solvation-induced processes affecting the relative stability of Zn2+(H2O)5(OH-), namely: (a) reduction of the Zn 4s character in solution states near the bottom of the conduction band; (b) hybridization between OH- orbitals and valence-band states of the solvent. Both effects contribute to hindering the OH- --> Zn2+ charge transfer, either by making it energetically unfavorable or by delocalizing the ligand charge density over several water molecules.  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