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 27 毫秒
1.
Restricted rotation about the naphthalenylcarbonyl bonds in the title compounds resulted in mixtures of cis and trans rotamers, the equilibrium and the rotational barriers depending on the substituents. For 2,7-dimethyl-1,8-di-(p-toluoyl)-naphthalene (1) ΔH° = 3.66 ± 0.14 kJ mol?1, ΔS° = 1.67 ± 0.63 J mol?1 K?1, ΔHct = 55.5 ± 1.3 kJ mol?1, ΔHct = 51.9 ± 1.3 kJ mol?1, ΔSct = ?41.3±4.1 J mol?1 K?1 and ΔSct = ?42.9±4.1 J mol?1 K?1. The rotation about the phenylcarbonyl bond requires ΔH = ?56.9±4.4 kJ mol?1 and ΔS = ?20.5±15.3 J mol?1 K?1 for the cis rotamer, and ΔH = 43.5Δ0.4 kJ mol?1 and ΔS =± ?22.4Δ1.3 J mol?1 K?1 for the trans rotamer. The role of electronic factors is likely to be virtually the same for both these rotamers but steric interaction between the two phenyl rings occurs in the cis rotamer only. Hence, the difference of the activation enthalpies obtained for the cis and trans rotamers, ΔΔH?1 = 13.4 kJ mol?1, provides a basis for the estimation of the role of steric factors in this rotation. For the tetracarboxylic acid 2 and its tetramethyl ester 3 the equilibrium is even more shifted towards the trans form because of enhanced steric and electrostatic interactions between the substituents in the cis form. The barriers for the rotation around the phenylcarbonyl bond and the cis-trans isomerization are lowered; an explanation for this result is presented.  相似文献   

2.
At room temperature and below, the proton NMR spectrum of N-(trideuteriomethyl)-2-cyanoaziridine consists of two superimposed ABC patterns assignable to two N-invertomers; a single time-averaged ABC pattern is observed at 158.9°C. The static parameters extracted from the spectra in the temperature range from –40.3 to 23.2°C and from the high-temperature spectrum permit the calculation of the thermodynamic quantities ΔH0 = ?475±20 cal mol?1 (?1.987 ± 0.084 kJ mol?1) and ΔS0 = 0.43±0.08 cal mol?1 K?1 (1.80±0.33 J mol?1 K?1) for the cis ? trans equilibrium. Bandshape analysis of the spectra broadened by non-mutual three-spin exchange in the temperature range from 39.4–137.8°C yields the activation parameters ΔHtc = 17.52±0.18 kcal mol?1 (73.30±0.75 kJ mol?1), ΔStc = ?2.08±0.50 cal mol?1 K?1 (?8.70±2.09 J mol?1 K?1) and ΔGtc (300 K) = 18.14±0.03 kcal mol?1 (75.90±0.13 kJ mol?1) for the transcis isomerization. An attempt is made to rationalize the observed entropy data in terms of the principles of statistical thermodynamics.  相似文献   

3.
Cyclohexane and piperidine ring reversal in 1-(3-pentyloxyphenylcarbamoyloxy)-2-dialkylaminocyclohexanes was investigated by 13C NMR. An unusually low conformational energy ΔG = 0.59 kJ mol?1 and activation parameters ΔG218 = 43.8 ± 0.4 kJ mol?1, ΔH = 48.9 ± 2.5 kJ mol?1 and ΔS = 23 ± 9 J mol?1 K?1 were found for the diequatorial to diaxial transition of the cyclohexane ring in the trans-pyrrolidinyl derivative. In the trans-piperidinyl derivative, ΔG222 = 44.7 ± 0.5 KJ mol?1, ΔH = 55.7 ± 6.3 kJ mol?1 and ΔS = 51 ± 21 J mol?1 K?1 was found for the piperidine ring reversal from the non-equivalence of the α-carbons.  相似文献   

4.
Cadmium thiourea reinickate undergoes two-stage thermal decomposition on heating. The DTG peak temperatures are 291 and 469°C and the corresponding DTA temperatures are 255 and 490°C. The kinetic parameters for the first stage decomposition are E* ≈ 120kJ mole?1; Z ≈ 1.2 × 108 cm3 mole?1 sec?1 and ΔS* ≈ ?95 J mole?1 K?1. For the second stage, E* ≈ 133 kJ mole?1; Z ≈ 6.1 × 105 cm?1 mole?1 sec?1 and ΔS* ≈ ?142 J mole?1 K?1.  相似文献   

5.
The solution structure and the aggregation behavior of (E)-2-lithio-1-(2-lithiophenyl)-1-phenylpent-1-ene ( 1 ) and (Z)-2-lithio-1-(2-lithiophenyl)ethene ( 2 ) were investigated by one- and two-dimensional 1H-, 13C-, and 6Li-NMR spectroscopy. In Et2O, both systems form dimers which show homonuclear scalar 6Li,6Li spin-spin coupling. In the case of 2 , extensive 6Li,1H coupling is observed. In tetrahdrofuran and in the presence of 2 mol of N,N,N′,N′-tetramethylethylylenediamine (tmeda), the dimeric structure of 1 coexists with a monomer. The activation parameters for intra-aggregate exchange in the dimers of 1 and 2 ( 1 (Et2O): ΔH≠ = 62.6 ± 13.9 kJ/mol, ΔS≠ = 5.8 ± 14.0 J/mol K, ΔG≠(263) = 61.1 kJ/mol; 2 (dimethoxyethane): ΔH≠ = 36.9 ± 6.5 kJ/mol, ΔS≠ = ?61 ± 25 J/mol K, ΔG≠(263) = 54.0 kJ/mol) and the thermodynamic parameters for the dimer-monomer equilibrium for 1 (ΔH°; = 26.7 ± 5.5 kJ/mol, ΔS° = 63 ± 27 J/mol K), where the monomer is favored at low temperature, were determined by dynamic NMR studies.  相似文献   

6.
The kinetics of the interactions between three sulfur‐containing ligands, thioglycolic acid, 2‐thiouracil, glutathione, and the title complex, have been studied spectrophotometrically in aqueous medium as a function of the concentrations of the ligands, temperature, and pH at constant ionic strength. The reactions follow a two‐step process in which the first step is ligand‐dependent and the second step is ligand‐independent chelation. Rate constants (k1 ~10?3 s?1 and k2 ~10?5 s?1) and activation parameters (for thioglycolic acid: ΔH1 = 22.4 ± 3.0 kJ mol?1, ΔS1 = ?220 ± 11 J K?1 mol?1, ΔH2 = 38.5 ± 1.3 kJ mol?1, ΔS2 = ?204 ± 4 J K?1 mol?1; for 2‐thiouracil: ΔH1 = 42.2 ± 2.0 kJ mol?1, ΔS1 = ?169 ± 6 J K?1 mol?1, ΔH2 = 66.1 ± 0.5 kJ mol?1, ΔS2 = ?124 ± 2 J K?1 mol?1; for glutathione: ΔH1 = 47.2 ± 1.7 kJ mol?1, ΔS1 = ?155 ± 5 J K?1mol?1, ΔH2 = 73.5 ± 1.1 kJ mol?1, ΔS2 = ?105 ± 3 J K?1 mol?1) were calculated. Based on the kinetic and activation parameters, an associative interchange mechanism is proposed for the interaction processes. The products of the reactions have been characterized from IR and ESI mass spectroscopic analysis. A rate law involving the outer sphere association complex formation has been established as   相似文献   

7.
The addition of thioacetic acid to unsaturated alcohols or acids was utilized to obtain mercaptoalkanols which were condensed with suitable carybonyl compounds to prepare 24 methyl-substituted 1,3-oxathianes. The 1H NMR spectra of the 1,3-oxathiane products were recorded at 60, 100 and/or 300 MHz and fully analysed. The results are best explained by a chair form which is completely staggered in the C-4? C-5? C-6 moiety ψ45 or (ψ56=60±1°). 1,3-Oxathianes having syn-axial 2,4- (and/or 2,6-) methyl-methyl interactions exist appreciably, if not exclusively, in twist forms. The vicinal coupling constants lead to the conformational free energies of axial methyl groups at C-4, ΔG°=7.4±0.4 kJ mol?1, and at C-5, ΔG°=3.7±0.3 kJ mol?1, in good agreement with previous estimates. They also show that both r-4,cis-5,trans-6- and r-4,trans-5,trans-6- trimethyl-1,3-oxathianes greatly favour the chiar form where the methyl group at C-4 is axial. The chair-twist energy parameters are reestimated at ΔH°CT 27.0 kJ mol?1, ΔS°CT 11.6J mol?1K?1, and ΔG°CT(298) 23.5 kJ mol?1 for a 2,5-twist form.  相似文献   

8.
2D 1H-1H EXSY NMR spectroscopy show that the free energy of activation ΔG in six 3-allyl-3-borabicyclo[3.3.1]nonane derivatives is significantly higher (72–86 kJ mol?1) than that in typical allylboranes (48–66 kJ mol?1). For the first member of the series, viz., 3-allyl-3-borabicyclo[3.3.1]nonane, the activation parameters of the permanent allylic rearrangement were also determined (ΔH = 82.7±3.4 kJ mol?1, ΔS = ?11.8±10.3 J mol?1 K?1, E A = 85.5±3.4 kJ mol?1, lnA = 29.2±1.2).  相似文献   

9.
Knudsen effusion studies of the sublimation of polycrystalline SnS, prepared by annealing and chemical vapor transport, have been performed employing vacuum micro-balance techniques in the temperature range 733–944 K and at pressures ranging from about 6 × 10?3 to 11 Pa.The third-law heats of sublimation and second-law entropy of reaction SnS(s) = SnS(g) were determined to be ΔH0298 = 220.4 ± 3.0 kJ mole? and ΔS0298 = 162.4 ± 4.5 J K?1 mole?1. From these data the standard heat of formation and absolute entropy of SnS(s) were calculated to be ?102.9 ± 4.0 kJ mole?1 and 79.9 ± 6.0 J K?1, respectively.  相似文献   

10.
Knudsen effusion studies of the sublimation of polycrystalline SnSe and SnSe2, prepared by annealing and chemical vapor transport reactions, respectively, have been carried out using vacuum microbalance techniques in the temperature ranges 736–967 K and 608–760 K, respectively. From experimental mass-loss data for the sublimation reaction SnSe(s) = SnSe(g), the recommended values for the heat of formation and absolute entropy of SnSe(s) were calculated to be ΔH°298,f = ?86.4 ± 9.9 kJ · mol?1 and S°298 = 89.0 ± 7.1 J · K?1 · mol?1. From mass-loss data for the decomposition reaction \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm SnSe}_{\rm 2} ({\rm s)} = {\rm SnSe(s)} + \frac{1}{{\rm x}}{\rm Se}_{\rm x} ({\rm g) (x} = 2 - 8) $\end{document}, the recommended values for the heat of formation and absolute entropy of SnSe2(s) were determined to be ΔH°298,f = ?118.1 ± 15.1 kJ · mol?1 and S°298 = 111.8 ± 11.8 J · K?1 mol?1.  相似文献   

11.
The kinetics of decomposition of an [Pect·MnVIO42?] intermediate complex have been investigated spectrophotometrically at various temperatures of 15–30°C and a constant ionic strength of 0.1 mol dm?3. The decomposition reaction was found to be first‐order in the intermediate concentration. The results showed that the rate of reaction was base‐catalyzed. The kinetic parameters have been evaluated and found to be ΔS = ? 190.06 ± 9.84 J mol?1 K?1, ΔH = 19.75 ± 0.57 kJ mol?1, and ΔG = 76.39 ± 3.50 kJ mol?1, respectively. A reaction mechanism consistent with the results is discussed. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 35: 67–72, 2003  相似文献   

12.

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   

13.
Iridium hexafluoride oxidizes ReF6 (via an ReF6+ salt) and at room temperatures IrF6, ReF6, ReF7 and (IrF5)4 are each present in the equilibrium mixture. From these and related findings: ΔH°(ReF6 → ReF6+ + e?) 1092 ± 27 kj mole?1(261 ± 6 kcal mole?1), and thermodynamic data are selected to yield ΔH°(ReF7(g) → ReF6+(g) + F?(g))=893 ± 33 kj mole?1(213 ± 8 kcal mole?1). From observations on the stability of IF6+BF4? and the lattice enthalpy evaluation for the salt, ΔH°(IF7(g) → IF6+(g) + F?(g))= 870 ± 24 kj mole?1(208 ± 6 kcal mole?1).  相似文献   

14.
The reversible dimerisation of o-phenylenedioxydimethylsilane (2,2-dimethyl-1,3,2-benzodioxasilole) has been studied by 1H NMR spectroscopy. The kinetics of this reaction can be described quantitatively by a bimolecular 10-ring formulation reaction and a monomolecular backreaction. The thermodynamic and kinetic parameters are: ΔH0 = ?43 kJ mol?1; ΔS0 = ?112 J mol?1 K?1; ΔG0298 = ?9.6 kJ mol?1; ΔH3298 = 57 kJ mol?1; ΔS3298 = ?129 J mol?1 K?1; ΔG3298 = 96 kJ mol?1; Ea = 60 kJ mol?1; A = 3.17 × 106 l mol?1 s?1. Remarkable is the low activation energy of formation of the ten-membered ring, considering that two SiO bonds have to be cleaved during the reaction. Transition states and possible structures of the ten-membered heterocycle are discussed.  相似文献   

15.
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   

16.
Ligand substitution kinetics for the reaction [PtIVMe3(X)(NN)]+NaY=[PtIVMe3(Y)(NN)]+NaX, where NN=bipy or phen, X=MeO, CH3COO, or HCOO, and Y=SCN or N3, has been studied in methanol at various temperatures. The kinetic parameters for the reaction are as follows. The reaction of [PtMe3(OMe)(phen)] with NaSCN: k1=36.1±10.0 s−1; ΔH1=65.9±14.2 kJ mol−1; ΔS1=6±47 J mol−1 K−1; k−2=0.0355±0.0034 s−1; ΔH−2=63.8±1.1 kJ mol−1; ΔS−2=−58.8±3.6 J mol−1 K−1; and k−1/k2=148±19. The reaction of [PtMe3(OAc)(bipy)] with NaN3: k1=26.2±0.1 s−1; ΔH1=60.5±6.6 kJ mol−1; ΔS1=−14±22 J mol−1K−1; k−2=0.134±0.081 s−1; ΔH−2=74.1±24.3 kJ mol−1; ΔS−2=−10±82 J mol−1K−1; and k−1/k2=0.479±0.012. The reaction of [PtMe3(OAc)(bipy)] with NaSCN: k1=26.4±0.3 s−1; ΔH1=59.6±6.7 kJ mol−1; ΔS1=−17±23 J mol−1K−1; k−2=0.174±0.200 s−1; ΔH−2=62.7±10.3 kJ mol−1; ΔS−2=−48±35 J mol−1K−1; and k−1/k2=1.01±0.08. The reaction of [PtMe3(OOCH)(bipy)] with NaN3: k1=36.8±0.3 s−1; ΔH1=66.4±4.7 kJ mol−1; ΔS1=7±16 J mol−1K−1; k−2=0.164±0.076 s−1; ΔH−2=47.0±18.1 kJ mol−1; ΔS−2=−101±61 J mol−1 K−1; and k−1/k2=5.90±0.18. The reaction of [PtMe3(OOCH)(bipy)] with NaSCN: k1 =33.5±0.2 s−1; ΔH1=58.0±0.4 kJ mol−1; ΔS1=−20.5±1.6 J mol−1 K−1; k−2=0.222±0.083 s−1; ΔH−2=54.9±6.3 kJ mol−1; ΔS−2=−73.0±21.3 J mol−1 K−1; and k−1/k2=12.0±0.3. Conditional pseudo-first-order rate constant k0 increased linearly with the concentration of NaY, while it decreased drastically with the concentration of NaX. Some plausible mechanisms were examined, and the following mechanism was proposed. [Note to reader: Please see article pdf to view this scheme.] © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 523–532, 1998  相似文献   

17.
Thermochemistry of Gaseous GeWO4 and GeW2O7 Mass spectrometric investigations with a Knudsen cell arrangement at temperatures between 1258 and 1383 K proved the existence of GeWO4 and GeW2O7 as component of the vaporphase over a mechanical mixture of GeO2 and WO2. Using the partial pressures heats of formation (2nd law calculation) and entropies (3rd law calculation) were computed; i. e. GeWO4: δH°1330 = ?149.8 kcal · mole?1, S°1330 = 129.9 cal K?1 mole?1, GeW2O7: δH°f,1330 = ?310.6 kcal mole?1, S°1330 = 190.0 cal K?1 mole?1. The standard heats of formation and entropies at 298 K, calculated with estimated Cp values are: GeWO4: δH°f,298 = ?181.6 kcal mole?1, S°298 = 85.0 cal K?1 mole?1; GeW2O7: δH°f,298 = ?365.8 kcal mole?1, S°298 = 112.1 cal K?1 mole?1. The thermochemical data of the GeWO4 and GeW2O7 molecules which also appear at chemical transport experiments [2] with GeO2 + WO2, are compared with known gaseous tungstates.  相似文献   

18.
E. M. F. of the Cell, Cd-Hg (2-phase)/CdAc2(m), Hg2Ac2(s)/Hg was measured at 20°, 25°, 30° and 40°C. The standard e. m. f. of the cell, Cd/CdAc3(m), Hg2Ac2(c)/Hg was evaluated as E°=1.1500?11.09×10?4T+1.06×10?8T2 The thermodynamic data of the reaction, Cd(c) + Hg2Ac2(c)=2Hg(l)+Cd++(aq)+2Ac?(aq) at 25°C were estimated as ΔF°=?42,139, ΔH°=?48,698 cal mole?1 and ΔS°=?22.0 cal deg?1 mole?1 at 25°C. The thermodynamic data for the formation of Hg2Ac2(s) were evaluated as ΔFf°=?202.3, ΔHf°=?154.5 Kcal mole?1 and S°=72.9 cal deg?1 mole?1. From measurements of the heats of solution of CdAc2·2H2O in aqueous solution, the relative partial molal enthalpies of cadmium acetate in aqueous solution were estimated.  相似文献   

19.
The thermal stability and kinetics of isothermal decomposition of carbamazepine were studied under isothermal conditions by thermogravimetry (TGA) and differential scanning calorimetry (DSC) at three heating rates. Particularly, transformation of crystal forms occurs at 153.75°C. The activation energy of this thermal decomposition process was calculated from the analysis of TG curves by Flynn-Wall-Ozawa, Doyle, distributed activation energy model, ?atava-?esták and Kissinger methods. There were two different stages of thermal decomposition process. For the first stage, E and logA [s?1] were determined to be 42.51 kJ mol?1 and 3.45, respectively. In the second stage, E and logA [s?1] were 47.75 kJ mol?1 and 3.80. The mechanism of thermal decomposition was Avrami-Erofeev (the reaction order, n = 1/3), with integral form G(α) = [?ln(1 ? α)]1/3 (α = ~0.1–0.8) in the first stage and Avrami-Erofeev (the reaction order, n = 1) with integral form G(α) = ?ln(1 ? α) (α = ~0.9–0.99) in the second stage. Moreover, ΔH , ΔS , ΔG values were 37.84 kJ mol?1, ?192.41 J mol?1 K?1, 146.32 kJ mol?1 and 42.68 kJ mol?1, ?186.41 J mol?1 K?1, 156.26 kJ mol?1 for the first and second stage, respectively.  相似文献   

20.
Nickel(II) chelate of 2–picolylamine has been studied spectrophotometrically in aqueous solution at 25°C and at an ionic strength of 0.3 M. The formation of pink color chelate was pH dependent, and the optimum pH range was between 7.0 to 8.5. Its mole ratio of ligand to nickel(II) ion was found to be 3 to 1 stoichiometry and the formation constant, logK, was determined as 13.31 ± 0.10. By using the wavelength 535 run, determination of trace amount of nickel(II) ion with the sensitivity of 5.28 τ/Cm2 was possible. Enthalpy and entropy changes characterizing the formation of the chelate have been calculated as follows: ΔG°=–8.15Kcal mole-1, ΔH°=–9.65 Kcal mole-1, ΔS°=28.5eu mole-1.  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