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 609 毫秒
1.
The aminolysis of Z‐thiophenyl methylacetates (C2H5C(O)SC6H4Z) with X‐benzylamines in acetonitrile has been investigated at 45°C. The reaction is found to proceed by a stepwise mechanism in which the rate‐determining step is the breakdown of the zwitterionic tetrahedral intermediate, T±, with possibly a hydrogen‐bonded four‐center‐type transition state. These mechanistic conclusions are drawn based on (i) the large magnitude of βX (= 1.2 ∼ 2.5) and βz (= −0.9 ∼ −1.5), (ii) the normal kinetic isotope effects (kH/kD ≅ 1.2) involving deuterated benzylamines (XC6H4CH2ND2), (iii) a large positive ρxz (= 2.4) and (iv) adherence to the reactivity‐selectivity principle in all cases. The extremely large βXnuc) values can be accounted for by the loss of a strong localized cationic charge on the N atom of benzylamines in the expulsion from the T±. The pKao (≥ 10.0) is high due to a large ratio of the expulsion rates of the amine (k−a) to thiophenolate (kb) (k−a/kb) from the T±. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 485–490, 2000  相似文献   

2.
The kinetics and mechanism of the reactions of O‐ethyl S‐(Z)aryl thiocarbonates with (X)benzylamines in acetonitrile at 45.0°C are studied. Relatively small values of βXnuc) = 0.6 ∼ 0.8 and βZlg) = −0.5 ∼ −0.7 together with a negative cross‐interaction constant ρXZ (= −0.47) and failure of the reactivity–selectivity principle (RSP) are interpreted to indicate a concerted mechanism. The normal kinetic isotope effects (kH/kD = 1.3 ∼ 1.8) involving deuterated benzylamine nucleophiles suggest a hydrogen‐bonded, four‐center‐type transition state. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 131–135, 2000  相似文献   

3.
The kinetics and mechanism of the nucleophilic substitution reactions of p‐chlorophenyl aryl chlorophosphates ( 2 ) with anilines are investigated in acetonitrile at 55°C. Relatively large magnitudes of ρX and βX values are indicative of a large degree of bond making in the TS. Smaller magnitudes of ρX (0.20 for X = H) and ρXY (?0.30) than those for the corresponding reactions with phenyl aryl chlorophosphates ( 1 ) (ρX = 0.54 for X = H and ρXY = ?1.31) are interpreted to indicate partial electron loss, or shunt, towards the electron acceptor equatorial ligand (p‐ClC6H4O‐) in the bipyramidal pentacoordinated transition state. The inverse secondary kinetic isotope effects (kH/kD = 0.64–0.87) involving deuterated aniline (ND2C6H4X) nucleophiles, and small ΔH? and large negative ΔS? are obtained. These results are consistent with a concerted nucleophilic substitution mechanism. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 632–637, 2002  相似文献   

4.
The kinetics of proton transfer between poly(A—AH) (partially protonated double-stranded polyadenylic acid) and CPR (chlorophenol red), and between poly(C—H—C) (partially protonated double-stranded polycytidylic acid) and the indicators CPR, BCP (bromocresol purple), and BCG (bromocresol green) have been investigated at 25°C and ionic strength 0.1 M (NaClO4) by the temperature-jump method. The acidic proton of poly(C—H—C) is engaged in a hydrogen bond (N3H+––––N3) which is believed to contribute to stabilizing the double-strand conformation, whereas the acidic proton of poly(A—A—H) does not form hydrogen bonds. The analysis of the dependence of the relaxation times on the concentrations of the reactants has enabled the evaluation of the rate constants for the direct proton transfer and for the protolysis paths. The rate constants for proton recombination with the deprotonated forms of the polynucleotides and the indicators are of the order of magnitude expected for diffusion controlled processes involving oppositely charged ions (k2=(0.2−1.6)×1010 M−1s−1). The direct proton transfer from poly(C—H—C) to BCG is thermodynamically disfavored and its rate constant, k1, is lower than k2 by about three orders of magnitude. The (thermodynamically favored) proton transfers from poly(A—A—H) to CPR and from poly(C—H—C) to CPR and BCP are characterized by similar values of k1. This result indicates that the hydrogen bonds in poly(C—H—C) are very weak and suggests that the stabilization of the double-stranded conformation of this polynucleotide could be ascribed to the large number of hydrogen bonds rather than to their specific strength. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 161–169, 1998.  相似文献   

5.
The aminolysis reactions of O-ethyl S-(Z-phenyl) dithiocarbonates (Z=p-CH3, H, p-Cl, and p-NO2) with anilines (AN) and N,N-dimethylanilines (DMA) in acetonitrile at 30.0°C are investigated. Relatively small values of βXnuc,0.4 ca. 0.7) and βZlg −0.1 ca. −0.4) for both ANs and DMAs, significantly large kH/kD values (1.1 ca. 1.9) involving deuterated anilines, and large negative ρXZ values for ANs (−0.56) are interpreted to indicate a concerted mechanism for both ANs and DMAs but with a hydrogen bonded four-center type transition state (TS) for ANs. The relative leaving ability, k(Z=p-NO2)/k(Z=p-CH3), is smaller for ANs than for DMAs, especially for a weaker nucleophile (1.9 and 4.7 for AN and DMA, respectively, with X=p-Cl). This suggests that the rate enhancement by the hydrogen-bond formation in the four-center type TS for AN is greater for a weaker nucleofuge (Z=p-CH3), especially when the nucleophile (X=p-Cl) is weaker. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 419–423,1998  相似文献   

6.
The fluorescence quenching coefficient (Ksv) and the quenching rate constant kq of ten para-substituted styrenes (1-Ys) have been measured and correlation-analyzed by both the dual-parameter equation (Eq. 1) with (ρxσx+ρ'σ') and the single-parameter equation (Eq. 2) with ρxσx. Ex-cellent results have been obtained for the correlation of KSV against (ρxσmb+ρ'σ'JJ) or (ρxσ+ +ρ'σ'JJ). Our results suggest that, possibly, there might be no need to use excited-state substituent constant for the fluorecence quenching process of excited states of styrenes.  相似文献   

7.
The solvolysis rates and products of several 1-substituted 2exo- and 2-endo-norbornyl p-toluenesulfonates 7 and 8 , respectively, have been determined. Hydrolyses of these epimeric tosylates yielded rearranged products in varying amounts, except when the substituent was COOCH3 or CN. The logarithms of the rate constants (log k) for the endo-series 8 correlated linearly with the corresponding inductive constants σ with a reaction constant ρI of ?1.24. On the other hand, log k values for the exo-series 7 appear to fit two regression lines, the first line (ρI = ?1.90) defined by the tosylates that ionize, with rearrangement, to the tertiary cations 11 , the second (ρI = ?1.86) by the tosylates 7 (R = H, COOCH3, and CN) that ionize to an asymmetrically bridged secondary cation 19 . These results confirm the unique participation of C(6) with a ρI of ?2.00 in the ionization of 2-exo-nor-bornyl tosylate.  相似文献   

8.
Reaction rates for the protodesilylation of trimethylphenylsilane and of [2,4,6-2H3]-trimethylphenylsilane by HCIO4, were measured in aqueous methanol (2:5, v/v) and the secondary deuterium isotope effect for the reaction was found to be kH/kD3 = 0.79. The magnitude of the observed isotope effect supports a mechanism in which the rate-determining step is the proton transfer from the hydronium ion to the silane to form a σ-intermediate.  相似文献   

9.
The catalytic activity of Pt nanoparticles (PtNPs) with different sizes and shapes was investigated in a photocatalytic hydrogen‐evolution system composed of the 9‐mesityl‐10‐methylacridinium ion (Acr+–Mes: photocatalyst) and dihydronicotinamide adenine dinucleotide (NADH: electron donor), based on rates of hydrogen evolution and electron transfer from one‐electron‐reduced species of Acr+–Mes (Acr.–Mes) to PtNPs. Cubic PtNPs with a diameter of (6.3±0.6) nm exhibited the maximum catalytic activity. The observed hydrogen‐evolution rate was virtually the same as the rate of electron transfer from Acr.–Mes to PtNPs. The rate constant of electron transfer (ket) increased linearly with increasing proton concentration. When H+ was replaced by D+, the inverse kinetic isotope effect was observed for the electron‐transfer rate constant (ket(H)/ket(D)=0.47). The linear dependence of ket on proton concentration together with the observed inverse kinetic isotope effect suggests that proton‐coupled electron transfer from Acr.–Mes to PtNPs to form the Pt? H bond is the rate‐determining step for catalytic hydrogen evolution. When FeNPs were used instead of PtNPs, hydrogen evolution was also observed, although the hydrogen‐evolution efficiency was significantly lower than that of PtNPs because of the much slower electron transfer from Acr.–Mes to FeNPs.  相似文献   

10.
The reactions of the title substrate (1) with a series of secondary alicyclic amines are subjected to a kinetic investigation in 44 wt% ethanol‐water, at 25.0°C, ionic strength 0.2 M (KCl). Under amine excess over the substrate, pseudo‐first‐order rate coefficients (kobs) are obtained. Plots of kobs against [NH], where NH is the free amine, are nonlinear upwards, except the reactions of piperidine, which show linear plots. According to the kinetic results and the analysis of products, a reaction scheme is proposed with two tetrahedral intermediates, one zwitterionic (T±) and another anionic (T), with a kinetically significant proton transfer from T± to an amine to yield T (k3 step). By nonlinear least‐squares fitting of an equation derived from the scheme to the experimental points, the rate microcoefficients involved in the reactions are determined. Comparison of the kinetics of the title reactions with the linear kobs vs. [NH] plots found in the same aminolysis of O‐ethyl 4‐nitrophenyl dithiocarbonate (2) in the same solvent shows that the rate coefficient for leaving group expulsion from T± (k2) is larger for 2 due to a stronger push by EtO than PhO. The k3 value is the same for both reactions since both proton transfers are diffusion controlled. Comparison of the title reactions with the same aminolysis of phenyl 4‐nitrophenyl thionocarbonate (3) in water indicates that (i) the k2 value is larger for the aminolysis of 1 due to the less basic nucleofuge involved and the small solvent effect on k2, (ii) the k3 value is smaller for the reactions of 1 due to the more viscous solvent, (iii) the rate coefficient for amine expulsion from T± (k−1) is larger for the aminolysis of 1 than that of 3 due to a solvent effect, and (iv) the value of the rate coefficient for amine attack (k1) is smaller for the aminolysis of 1 in aqueous ethanol, which can be explained by a predominant solvent effect relative to the electron‐withdrawing effect from the nucleofuge. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 839–845, 1999  相似文献   

11.
The expression of pseudo-second-order rate constants (k X) for cationic nanoparticle (CN) [CTABr/NaX/H2O, X = Br, Cl, CTABr = cetyltrimethylammonium bromide] catalyzed piperidinolysis-ionized phenyl salicylate (PSa), at constant [CTABr]T, 0.1 M piperidine (Pip), and 35°C, were calculated from the relationship: k obs = (k 0 + k X[NaX])/(1 + K X/S[NaX]), in which k 0, k X, and K X/S are constant kinetic parameters and k obs represents the pseudo-first-order rate constant for Pip reaction with phenyl salicylate ion in the presence of CN. The source of the large catalytic effect of CN catalyst was shown to be due to the transfer of PSa from pseudo-phase of the CNs to the bulk aqueous phase through X/PSa ion exchange at the surface of the CNs.  相似文献   

12.
The cationic [RuCl(PNNP)]+ catalysts containing tetradentate ligands with a P2N2 donor set (PNNP) show strong electronic effects in the cyclopropanation of para‐substituted styrenes. The reactivity trend confirms that the carbene transfer to the olefin has electrophilic character. Linear free‐energy relationships are observed for the relative reactivity, the cis/trans selectivity, and for the enantioselectivity (of the cis‐cyclopropane). The linear correlation between log (kX/kH) and σ shows a large value of ρ (−2.4), which is indicative of significant charge buildup in the transition state of the carbene transfer to the olefin. All the relevant parameters (reactivity, diastereoselectivity, and enantioselectivity) increase with the increasing electron density at the C=C bond. To define the scope of the [RuCl(PNNP)]+ catalysts, 1‐ and 2‐substituted styrenes, and oct‐1‐ene were also investigated.  相似文献   

13.
Isotope effects, general acid catalysis, and relative reactivities show that proton transfer to one of the unsaturated carbon atoms is rate determining for the acidolysis of unsaturated alkylmercuric halides. For compounds, R1R2C?CHHgX, substitution of CH3 for H at R1 or R2 leads to an acceleration of a factor of ~ 30. This relatively small acceleration, the relative facility of the reactions, and the magnitude of the Br? catalytic terms, suggests an olefin–mercuric halide complex as the product of the rate-determining step, rather than a simple carbonium ion. The Brøonsted catalysis law is obeyed with a variety of carboxylic acids, giving an ∝ of 0.69 ± 0.04, but acids of other structures give substantially deviant catalytic coefficients, in a pattern similar to that generated by other A-SE2 reactions. The acetic acid catalytic coefficient is larger by a factor of 102 than that predicted if it were due to specific hydronium ion–general base catalysis instead of true general acid catalysis. The overall solvent isotope effect, kH/kD, is 2.55 ± 0.10. The competitive isotope effect, κHD, is 6.84 ± 0.06. Taken with a model in which the proton is transferred directly from the H3O+ unit of the aquated proton to the substrate, these are sufficient to successfully predict the rate at all intermediate isotopic compositions.  相似文献   

14.
Abstract

Molten copper-tin alloys have been studied by X-ray diffraction, using a focusing theta-theta diffraetometer and Mo-Kα radiation (monochromator in the diffracted beam). Five alloys with 20, 35, 45, 55 and 78 atomic percent Sn, and pure Cu and Sn were measured at temperatures about 20 °C above the liquidus, and at 1100 °C. The total interference functions I(K), where K = 4π sin θ/δ, were obtained from the observed scattered intensities Ia(K) per atom and the theoretical atomic scattering factors. Splitting of the first peak in I(K) has been observed in the Cu-55 at% Sn alloy at the liquidus temperature.

The partial interference functions Itj(K) at the liquidus temperature and at 1100°C were evaluated (assuming that they are independent of atomic concentration) using the five total I(K) of the alloys. The functions Iij(K) are in reasonable agreement with those obtained by Enderby, North and Egelstaff from neutron diffraction data of a Cu-45 at % Sn alloy.

The reduced partial distribution functions Gij(r) = 4πρ0 r{gij(r) ? 1} and the probability functions gij(r) = ρ ij(r)/cj ρ0, where ρ ij(r) is the number of j-type atoms per unit volume at the distance r from an i-type atom, cj is the atomic fraction of j-type atoms and ρ0 is the average atomic density, have been evaluated by Fourier transformation of {Iij(K) ? 1} K.

The electrical resistivities ρ R of the alloys, calculated with the Faber-Ziman equation using the measured Iij(K) and Animalu-Heine pseudo-potential elements Ui(K), are in good agreement with the experimental values of Roll and Motz. Assuming that Ui (2k F) is independent of the values of the Fermi diameter 2k F of the alloys, the concentration dependence of (3 - XR, where X is the thermoelectric parameter measured by Enderby and Howe, is well reproduced when using the X-ray values of Iij (2k F).  相似文献   

15.
Oxidation of meta- and para-substituted benzylamines by cetyltrimethylammonium permanganate (CTAP) to the corresponding aldimines is first order with respect to both the amine and CTAP. Oxidation of deuteriated benzylamine (PhCD2NH2) exhibited the presence of a substantial kinetic isotope effect (k H /k D = 5.60 at 293 K). This confirmed the cleavage of an α-C-H bond in the rate-determining step. Correlation analyses of the rates of oxidation of 19 monosubstituted benzylamines were performed with various single and multiparametric equations. The rates of the oxidation showed excellent correlations in terms of Yukawa—Tsuno and Brown’s equations. The polar reaction constants are negative. The oxidation exhibited an extensive cross-conjugation, in the transition state, between the electron-donating substituents and the reaction centre. A mechanism involving a hydride-ion transfer from the amine to CTAP in the rate-determining step has been proposed.  相似文献   

16.
The oxidation of primary alcohols by sodium N-chloroethylcarbamate in acid solution, results in the formation of corresponding aldehydes. The reaction is first order with respect to the oxidant and alcohol. The rate increases with an increase in acidity. The oxidation of α,α-dideuterioethanol exhibited a primary kinetic isotope, kH/kD = 2.11 at 298 K. The value of solvent isotope effect k(H2O)/k(D2O) = 2.23 at 298 K. Addition of ethyl carbamate does not affect the rate. (EtOC(OH)NHCl)+ has been postulated as the reactive species. Plots of (log k2 + Ho) against (Ho + log[H+]) are linear with the slope, ?, having values from 1.78–1.87. This suggested a proton abstraction by water in the rate-determining step. The rates of oxidation of alcohols bearing both electron-withdrawing and electron-donating groups are more than that of methanol. A concerted mechanism involving transfer of a hydride ion from the C? H bond of the alcohol tothe oxidant and removal of a proton from the O? H group by a water molecule has been proposed.  相似文献   

17.
The local density approximation (LDA) to the exchange potential Vx( r ), namely the ρ1/3 electron gas form, was already transcended in Slater's 1951 paper. Here, using Dirac's 1930 form for the exchange energy density ? x( r ), the Slater (Sl) nonlocal exchange potential V( r ) is defined by 2? x( r )/ρ( r ). In spherical atomic ions, say the Be or Ne‐like series, this form V( r ) already has the correct behavior in both r → 0 and r → ∞ limits when known properties of the exchange energy density ? x( r ) and the ground‐state electron density ρ( r ) are invoked. As examples, some emphasis will first be given to the use of the so‐called 1/Z expansion in such spherical atomic ions, for which analytic results can be obtained for both ? x( r ) and ρ( r ) as the atomic number Z becomes large. The usefulness of the 1/Z expansion is directly demonstrated for the U atomic ion with 18 electrons by comparison with the optimized effective potential prediction. A rather general integral equation for the exchange potential is then proposed. Finally, without appeal to large Z, two‐level systems are considered, with specific reference to the Be atom and to the LiH molecule. In all cases treated, the Slater potential V( r ) is a valuable starting point, even though it needs appreciable quantitative corrections reflecting directly atomic shell structure. © 2004 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

18.
The kinetics of addition of a number of ortho‐, meta‐, and para‐substituted benzylamines to benzylidenemalononitrile (BMN) in acetonitrile have been studied. The reaction is first‐order with respect to BMN. The order with respect to the amine is more than one. It has been shown that the reaction followed two mechanistic pathways, uncatalyzed and catalyzed by the amine. The enthalpy of activation for the catalyzed path is negative indicating the presence of a preequilibrium (k1, k−1) leading to the formation of a zwitterion. The values of rate constant, k1, for the nucleophilic attack have been determined for twenty‐eight benzylamines. The rate constant, k1 was subjected to correlation analyses using various single‐ and multi‐parametric equations. The best correlation is obtained in terms of Charton's LDR and LDRS equations. The polar regression coefficients are negative indicating the formation of a cationic species in the transition state. The reaction is subject to steric hindrance by ortho‐substituents. © 1999 John Wiley & Sons, Inc., Int J Chem Kinet 31: 245–252, 1999  相似文献   

19.
Various p-substituted benzyl p-hydroxyphenyl methyl sulfonium salts ( 2 ) were synthesized and their initiator activities were evaluated in bulk polymerization of glycidyl phenyl ether (PGE). The order of the activity was found to be 2b (X = CH3) > 2a (X = H) ≈ 2c (X = Cl) > 2d (X = NO2), indicating that the introduction of an electron-donating group enhanced the activity. In Hammett's plots, the logarithm of the ratio of the polymerization rates (log kx/kH) was correlated with σ+ρ better than with σp and a negative ρ+ value (-1.18) was obtained. Reaction of 2a with benzyl mercaptan mainly gave dibenzyl sulfide and p-hydroxyphenyl methyl sulfide. The obtained results seemed to demonstrate that the OH group of the aryl group yielded no proton as initiator for the polymerization, whereas the benzyl group caused the polymerization, which was initiated by the corresponding benzyl cation formed by C? S bond cleavage. © 1993 John Wiley & Sons, Inc.  相似文献   

20.
Temperature dependences of the relative reactivity of potassium aryloxides XC6H4O?K+ toward 4‐nitrophenyl ( 1 ), 3‐nitrophenyl ( 2 ), 4‐chlorophenyl ( 3 ), and phenyl ( 4 ) benzoates in dimethylformamide (DMF) were studied using the competitive reactions technique. The rate constants kX for the reactions of 1 with potassium 4‐cyanophenoxide, 2 with potassium 3‐bromophenoxide, 3 with potassium 3‐bromo‐, 4‐bromo‐, and unsubstituted phenoxides, 4 with potassium 4‐methoxy‐ and 3‐methylphenoxides were measured at 25°C. Correlation analysis of the relative rate constants kX/kH(3‐Me) and differences in the activation parameters (??Н and ??S) of competitive reactions revealed the existence of six isokinetic series. We investigated the substituent effect of X on the activation parameters for each isokinetic series and concluded that the reactions of aryl benzoates PhCO2C6H4Y with potassium aryloxides in DMF proceed via a four‐step mechanism. The large ρ0(Y) and ρXY values at 25°C obtained for the reactions of 1–3 with potassium aryloxides with an electron‐donating substituent refer to the rate‐determining formation of the spiro‐σ‐complex. The Hammett plots for the reactions of 1 and 2 exhibit a downward curvature, causing the motion of the transition state for the rate‐determining step according to a Hammond effect as the substituent in aryloxide changes from electron‐donating to electron‐withdrawing. Analysis of data in the terms of two‐dimensional reaction coordinate diagrams leads to the conclusion that significant anti‐Hammond effects arise in the cases of ortho‐substituted and unsubstituted substrates. It was shown that the isokinetic and compensation effects observed for the reactions of aryl benzoates with potassium aryloxides in DMF can be interpreted in the terms of the electrostatic bonding between the reaction centers.  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