首页 | 官方网站   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.
Kinetics of the substitution reaction of solvent molecule in uranyl(VI) Schiff base complexes by tri‐n‐butylposphine as the entering nucleophile in acetonitrile at 10–40°C was studied spectrophotometrically. The second‐order rate constants for the substitution reaction of the solvent molecule were found to be (8.8 ± 0.5) × 10?3, (5.3 ± 0.2) × 10?3, (7.5 ± 0.3) × 10?3, (6.1 ± 0.3) × 10?3, (13.5 ± 1.6) × 10?3, (13.2 ± 0.9) × 10?3, (52.9 ± 0.2) × 10?3, and (88.1 ± 0.6) × 10?3 M?1 s?1 at 40°C for [UO2(Schiff base)(CH3CN)], where Schiff base = L1–L8, respectively. In a temperature dependence study, the activation parameters ΔH# and ΔS# for the reaction of uranyl complexes with PBu3 were determined. From the linear rate dependence on the concentration of PBu3, the span of k2 values and the large negative values of the activation entropy, an associative (A) mechanism is deduced for the solvent substitution. By comparing the second‐order rate constants k2, it was concluded that the steric and the electronic properties of the complexes were important for the rate of the reactions.  相似文献   

2.
The complex [Fe(imox) (Nmim)2], where imox is a planar bis(iminooxime) macrocyclic ligand and Nmim=N-methyl imidazole, undergoes substitution reactions in the presence of 2,2′bipyridine (bipy), leading to a series of intermediate mixed complexes. The forward and reverse rate constants for the substitution of the Nmim ligand by water are k1 = 0.23 s?1 and k?1=62 M?1s?1, respectively. The binding of a bipy to the [Fe(imox) (Nmim) (H2O)]+ complex proceeds according to k2 = 4.7×10?2 M?1 s?1 and k?2 = 3.6 × 10?4 s?1, yielding [Fe(imox) (bipy) (Nmim)]+, which eliminates Nmim according to k3 = 1.6 × 10?4 s?1. Further substitution in the [Fe(imox) (bipy)]+ complex with bipy takes place very slowly leading to the [Fe(imox) (bipy)2]+ and [Fe(bipy)3]2+ complexes. © 1993 John Wiley & Sons, Inc.  相似文献   

3.

Ligand substitution of trans-[CoIII(en)2(Me)H2O]2+ was studied for pyrazole, 1,2,4-triazole and N-acetylimidazole as entering nucleophiles. These displace the coordinated H2O molecule trans to the methyl group to form trans-[Co(en)2(Me)azole]. Stability constants at 18°C for the substitution of H2O by pyrazole, 1,2,4-triazole and N-acetylimidazole are 0.7 ± 0.1, 13.8 ± 1.4 and 1.7 ± 0.2 M?1, respectively. Second order rate constants at the same temperature for the reaction of trans-[CoIII(en)2(Me)H2O]2+ with pyrazole, 1,2,4-triazole and N-acetylimidazole are 161 ± 12, 212 ± 11 and 12.9 ± 1.6 M?1 s?1, respectively. Activation parameters (ΔH, ΔS) are 67 ± 6 kJ mol?1, + 27 ± 19 J K?1 mol?1; 59 ± 2 kJ mol?1, + 1 ± 6 J K?1 mol?1 and 72 ± 4 kJ mol?1, + 23 ± 14 J K?1 mol?1 for reactions with pyrazole, 1,2,4-triazole and N-acetylimidazole, respectively. Substitution of coordinated H2O by azoles follows an Id mechanism.  相似文献   

4.
A series of dinuclear platinumII complexes of the type [{trans‐Pt(H2O)(NH3)2}2‐NH2(CH2)nH2N]4+ (where n = 2, 3, 4, and 6) were synthesized to investigate the influence of the bridging diamine linker on the reactivity of the platinum centers. The pKa values were determined, and the rates of substitution of the aqua moieties by a series of neutral nucleophiles viz. thiourea, 1,3‐dimethyl‐2‐thiourea, and 1,1,3,3‐tetramethyl‐2‐thiourea were studied as a function of concentration and temperature. All reactions studied gave excellent fits to a single exponential and obeyed the simple rate law, kobs=k2[Nu]. Negative activation entropies support an associative mode of substitution. The results obtained suggest that the rate of substitution is definitely influenced by the length of the diamine chain, with the rate of substitution decreasing as the length of the diamine chain increases. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 38: 202–210, 2006  相似文献   

5.
The partial isotope substitution for the change of geometrical parameters, interaction energies, and nuclear magnetic shielding tensors (σ) of dihydrogen‐bonded NH3X+···YBeH (X, Y = H, D, and T) systems is analyzed. Based on the theoretical calculation, the distance between heavy atoms RN···Be of NH3H+···DBeH is clearly found to be shorter than that in NH3D+···HBeH. Such apparently paradoxical geometrical isotope effect (GIE) on RN···Be is revealed by the cooperative effect of two kinds of (1) primary covalent‐bonded GIE and (2) secondary dihydrogen‐bonded one. We have demonstrated that (1) the covalent bond lengths become shorter by heavier isotope‐substitution and (2) the dihydrogen‐bonded distance RX···Y becomes shorter by heavier Y and lighter X isotope‐substitution due to the difference of electronic structure reflected by the nuclear distribution. We have also found that interaction energy of NH3H+···DBeH is stronger than that of NH3D+···HBeH and isotopic deshielding effect of magnetic shielding becomes large in lighter isotope. © 2013 Wiley Periodicals, Inc.  相似文献   

6.
The kinetics of the homolytic substitution of several trialkyltin iodides by iodine atoms are presented. Rate constants have been determined at three different temperatures and the following activation parameters calculated: A, Ea, and ΔS°. The observation that the activation energy, ΔG, is related to the driving force of the ion-pair formation, leads to the conclusion that the charge-transfer model is a valid approach for substitution in the reaction between R3SnI compounds and iodine atoms.  相似文献   

7.
Reaction rates for the structural isomerization of 1,1,2,2‐tetramethylcyclopropane to 2,4‐dimethyl‐2‐pentene have been measured over a wide temperature range, 672–750 K in a static reactor and 1000–1120 K in a single‐pulse shock tube. The combined data from the two temperature regions give Arrhenius parameters Ea=64.7 (±0.5) kcal/mol and log10(A, s?1) = 15.47 (±0.13). These values lie at the upper end of the ranges of Ea and log A values (62.2–64.7 kcal/mol and 14.82–15.55, respectively) obtained from three previous experimental studies, each of which covered a narrower temperature range. The previously noted trend toward lower Ea values for structural isomerization of methylcyclopropanes as methyl substitution increases extends only through the dimethylcyclopropanes (1,1‐ and 1,2‐); Ea then appears to increase with further methyl substitution. In contrast, the pre‐exponential factors for isomerization of cyclopropane and all of the methylcyclopropanes through tetramethylcyclopropane lie within ±0.3 of log10(A, s?1) = 15.2 and show no particular trend with increasing substitution. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 38: 483–488, 2006  相似文献   

8.
The rate for the substitution reaction of Co(edta)? with ethylenediamine was greatly enhanced by the presence of an excess of Co(II) ion in solution. The rate constant is (13±2) M?-sec?1 at μi=0.10M LiClO4, pH=11.1, [en]=0.10M and T=25°C. The mechanism for the reaction is discussed on the basis of the Marcus theory for outer-sphere processes corrected for electrostatic effects. This catalytic effect was not observed when the Co(II) was present in small amount due to the stability of the Co(edta)?2 complex toward substitution. The rate constant for direct substitution of Co(edta)? under the same conditions has also been measured and the value is (3.66±0.40)×10?4sec?.  相似文献   

9.
10.
A tetragonal 123 phase with the composition close to CeLa2 { Cu 2 2+ } [Mg2+]O8 (the braces indicate the Cu(2) positions; the brackets indicate the Cu(1) positions) with the parameters a = b = 0.3909(3) nm, c = 1.6591(8) nm was prepared at 860°C under an oxygen atmosphere with an elevated oxygen pressure. When the lanthanum-for-barium substitution was incomplete, the resulting 123 phase had the composition close to CeLa1.7Ba0.3{ Cu 1.7 2+ } [Mg]O8 with the unit cell parameters a = b = 0.3868(3) nm, c = 1.6578(8) nm that contains Cu3+ in the Cu(2) positions. The partial substitution of barium for lanthanum (the melting point of barium oxide is almost 500°C lower that of the lanthanum oxide) appreciably facilitated the synthesis: the 123 phase in this sample was more than 90%. The existence of Cu3+ in the Cu(2) positions enhanced the electrical conductivity of the sample.  相似文献   

11.
New high-conductance potassium-cation solid electrolytes based on potassium aluminate are synthesized by means of partial substitution of five-charged phosphorus cations for three-charged aluminum cations and investigated. The maximum conductivity in the system K2 − 2x Al2 − x P x O4 is found to equal 5 × 10−3 S cm−1 at 200°C and ∼1 S cm−1 at 700°C, which is one of the best values for potassium solid electrolytes. The principal factors responsible for the high conductance are the stabilization of a high-temperature form of potassium aluminate and the formation of additional vacancies in the potassium sublattice, which occurs during the substitution process Al3+ → P5+ + 2V K . __________ Translated from Elektrokhimiya, Vol. 41, No. 12, 2005, pp. 1501–1504. Original Russian Text Copyright ? 2005 by Burmakin, Shekhtman.  相似文献   

12.
The kinetics of the interaction of L ‐asparagine with [Pt(ethylenediamine)(H2O)2]2+ have been studied spectrophotometrically as a function of [Pt(ethylenediamine)(H2O)22+], [L ‐asparagine], and temperature at pH 4.0, where the substrate complex exists predominantly as the diaqua species and L ‐asparagine as the zwitterion. The substitution reaction shows two consecutive steps: the first step is the ligand‐assisted anation and the second one is the chelation step. Activation parameters for both the steps have been calculated using Eyring equation. The low ΔH1 (43.59 ± 0.96 kJ mol?1) and large negative values of ΔS1 (?116.98 ± 2.9 J K?1 mol?1) as well as ΔH2 (33.78 ± 0.51 kJ mol?1) and ΔS2 (?221.43 ± 1.57 J K?1 mol?1) indicate an associative mode of activation for both the aqua ligand substitution processes. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 252–259, 2003  相似文献   

13.
The rate of substitution of the chloride and aqua moieties from the platinum(II)-amine complexes, viz. [Pt(dien)Cl]Cl(Pt1-Chloro) and [Pt(en)(NH3)Cl]Cl (Pt2-Chloro) and their corresponding aqua analogues, viz. [Pt(dien)(OH2)] (ClO4)2 (Pt1-Aqua) and [Pt(en)(NH3)(OH2)](ClO4)2 (Pt2-Aqua), by a series of neutral and anionic nucleophiles,viz. thiourea (TU), 1,3-dimethyl-2-thiourea (DMTU), 1,1,3,3-tetramethyl-2-thiourea (TMTU), iodide (I) and thiocyanate (SCN), was determined under pseudo first-order conditions as a function of concentration and temperature using UV/Visible spectrophotometry and standard stopped-flow techniques. The observed pseudo first-order rate constants for the substitution reactions obeyed the simple rate law k obs = k 2[Nucleophile]. Second-order kinetics and negative activation entropies, ca. −93 J K−1 mol−1 and −71 J K−1 mol−1, for the chloro and aqua complexes respectively, support an associative mode of activation. The rate of substitution of both the chloro and aqua moieties are observed to decrease with an increase in the steric bulk of the neutral nucleophiles, whilst rate of substitution by SCN was observed to be faster than that of I, in correlation with the observed nucleophilicities of the two nucleophiles. A comparison of the second-order rate constants, k 2, at 298 K, obtained for the substitution reactions of Pt1and Pt2 shows that an increase in chelation in moving from Pt2 to Pt1 results in a corresponding increase in the reactivity, by a factor of ca. 3, (28.31 ± 0.15 and 8.02 ± 0.13 m −1 s−1 for Pt1 and Pt2 respectively, in the case of substitution of the aqua species by TU). Computational analysis of the chloro complexes, viz. Pt1-Chloro, Pt2-Chloro and [Pt(NH3)3Cl]Cl (Pt3) support this conclusion by demonstrating that the Pt–N bond trans to the leaving group is shortened and that the Pt–Cl bond is lengthened when chelation is increased from Pt3 to Pt1. Consequently, these results suggest that the increase in reactivity of Pt1 over Pt2, promoted by increased chelation, is as a result of ground state destabilization.  相似文献   

14.
The quenching rate constants of O2(1Δg) with n-butylamine, diethylamine, dipropylamine, dibutylamine, and tripropylamine have been determined in a discharge flow system. The rate constants are found to be (1.6 ± 0.2) × 103, (8.5 ± 0.6) × 104, (9.8 ± 0.5) × 104, (2.1 ± 0.1) × 105, and (8.6 ± 0.5) × 105 1 mol?1 s?1, respectively. The rate constants are found to increase in the order, tertiary amine → secondary amine → primary amine. The “inductive effect” of alkyl substitution is also found to increase the rate constant in a given series of amines.  相似文献   

15.
An approach towards precision NMR measurements of four‐bond deuterium isotope effects on the chemical shifts of backbone amide nitrogen nuclei in proteins is described. Three types of four‐bond 15 N deuterium isotope effects are distinguished depending on the site of proton‐to‐deuterium substitution: 4ΔN(Ni‐1D), 4ΔN(Ni+1D) and 4ΔN(Cβ,i‐1D). All the three types of isotope shifts are quantified in the (partially) deuterated protein ubiquitin. The 4ΔN(Ni+1D) and 4ΔN(Cβ,i‐1D) effects are by far the largest in magnitude and vary between 16 and 75 ppb and ?18 and 46 ppb, respectively. A semi‐quantitative correlation between experimental 4ΔN(Ni+1D) and 4ΔN(Cβ,i‐1D) values and the distances between nitrogen nuclei and the sites of 1H‐to‐D substitution is noted. The largest isotope shifts in both cases correspond to the shortest inter‐nuclear distances. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

16.
The kinetics of the substitution reaction of uranyl Schiff base complexes with tributylphosphine was studied spectrophotometrically in acetonitrile. Uranyl complexes have a pentagonal bipyramidal structure with a trans‐UO2 moiety at the axial positions. In uranyl tetradentate Schiff base complexes, the fifth position of the equatorial plane is occupied by the solvent molecule, which weakly coordinates to the U center. In a substitution reaction, tributylphosphine can easily replace the solvent molecule. By considering the excellent linearity of kobs versus the molar concentration of tributylphosphine, the large negative values of Δ S#, and the small values of Δ H#, an associative (A) mechanism has been suggested. By comparing the rate constants (k2) and the activation parameters, it is obvious that two parameters are effective in the rate of substitution reactions; The first parameter is the steric effect that the rate of reaction has been decreased by increasing this factor, and the other parameter is the electronic property that the electron‐withdrawing group leads to increase the rate of reaction and the electron donor group decreases it. © 2013 Wiley Periodicals, Inc. Int J Chem Kinet 45: 168–174, 2013  相似文献   

17.
Cloud point measurements were made for binary systems of polystyrene (PS) + methyl acetate (MA) and polystyrene (PS) + selectively deuterated MA: CD3COOCH3 and CH3COOCD3 with three PS samples of Mw = 4.0 × 105, 2.0 × 106, and 13.2 × 106. All systems are characterized by the phase diagrams with upper and lower critical temperature. H/D isotope effects on miscibility for both selectively deuterated acetates are very large and appeared to be independent of the site of deuterium substitution. © 2009 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 47:2140–2143, 2009  相似文献   

18.
Substitution reactions of a Cl ligand in [SnCl2(tpp)] (tpp=5,10,15,20‐tetraphenyl‐21H,23H‐porphinato(2−)) by five organic bases i.e., butylamine (BuNH2), sec‐butylamine (sBuNH2), tert‐butylamine (tBuNH2), dibutylamine (Bu2NH), and tributylamine (Bu3N), as entering nucleophile in dimethylformamide at I=0.1M (NaNO3) and 30–55° were studied. The second‐order rate constants for the substitution of a Cl ligand were found to be (36.86±1.14)⋅10−3, (32.91±0.79)⋅10−3, (22.21±0.58)⋅10−3, (19.09±0.66)⋅10−3, and (1.36±0.08)⋅10−3 M −1s−1 at 40° for BuNH2, tBuNH2, sBuNH2, Bu2NH, and Bu3N, respectively. In a temperature‐dependence study, the activation parameters ΔH and ΔS for the reaction of [SnCl2(tpp)] with the organic bases were determined as 38.61±4.79 kJ mol−1 and −150.40±15.46 J K−1mol−1 for BuNH2, 40.95±4.79 kJ mol−1 and −143.75±15.46 J K−1mol−1 for tBuNH2, 30.88±2.43 kJ mol−1 and −179.00±7.82 J K−1mol−1 for sBuNH2, 26.56±2.97 kJ mol−1 and −194.05±9.39 J K−1mol−1 for Bu2NH, and 39.37±2.25 kJ mol−1 and −174.68±7.07 J K−1 mol−1 for Bu3N. From the linear rate dependence on the concentration of the bases, the span of k2 values, and the large negative values of the activation entropy, an associative (A) mechanism is deduced for the ligand substitution.  相似文献   

19.
The kinetics of the interaction of adenosine with cis‐[Pt(cis‐dach)(OH2)2]2+ (dach = diaminocyclohexane) was studied spectrophotometrically as a function of [cis‐[Pt(cis‐dach)(OH2)2]2+], [adenosine], and temperature at a particular pH (4.0), where the substrate complex exists predominantly as the diaqua species and the ligand adenosine exists as a neutral molecule. The substitution reaction shows two consecutive steps: the first is the ligand‐assisted anation followed by a chelation step. The activation parameters for both the steps have been evaluated using Eyring equation. The low negative value of ΔH1 (43.1 ± 1.3 kJ mol?1) and the large negative value of ΔS1 (?177 ± 4 J K?1 mol?1) along with ΔH2 (47.9 ± 1.8 kJ mol?1) and ΔS2 (?181 ± 6 J K?1 mol?1) indicate an associative mode of activation for both the aqua ligand substitution processes. The kinetic study was substantiated by infrared and electrospray ionization mass spectroscopic analysis. © 2011 Wiley Peiodicals, Inc. Int J Chem Kinet 43: 219–229, 2011  相似文献   

20.
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   

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-23

京公网安备 11010802026262号