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.
Excess molar enthalpies HmE and excess molar volumes VmE have been measured for xC3H7NO2 + (1 ? x)c-C6H12 at 298.15 and 318.15 K; +(1 ? x)CCl4 at 298.15 and 318.15 K; +(1 ? x)C6H6 at 298.15 and 318.15 K; +(1 ? x)C6H14 (VmE only) at 298.15 K; +(1 ? x)p-C6H4(CH3)2 at 298.15 K; and for xCH3CH(NO2)CH3 + (1 ? x)c-C6H12 at 298.15 and 318.15 K; +(1 ? x)CCl4 at 298.15 and 318.15 K; +(1 ? x)C6H6 at 298.15 K; +(1 ? x)C6H14 at 298.15 K; +(1 ? x)(CH3)2CHCH(CH3)2 for HmE at 318.15 K and for VmE at 298.15 K; and +(1 ? x)C16H34 at 298.15 K. The HmE′s were determined with an isothermal dilution calorimeter and the VmE′s with a continuous-dilution dilatometer. Particular attention was paid to the region dilute in nitroalkane. In general HmE is large and positive for (a nitropropane + an alkane), less positive for (a nitropropane + tetrachloromethane), and small for (a nitropropane + benzene) and for (a nitropropane + 1,4-dimethylbenzene). The mixture with hexadecane shows phase separation. VmE is large and positive for (1-nitropropane + cyclohexane), less positive for (1-nitropropane + hexane), and S-shaped for (1-nitropropane + tetrachloromethane) with negative values in the 1-nitropropane-rich region. For (1-nitropropane + benzene) and for (1-nitropropane + 1,4-dimethylbenzene) VmE is negative. For mixtures with 2-nitropropane the results are similar except that for benzene VmE is S-shaped with positive values in the 2-nitropropane-rich region.  相似文献   

2.
Vapour pressures, excess enthalpies, and densities for {(1?x)C6H14 + xCS2} {(1?x)C10H22 + xCS2}, {(1?x)C13H28 + xCS2}, and {(1?x)C16H34 + xCS2} have been measured at 298.15 K. It was found that HmE and VmE increase as chain length increases while GmE diminishes, becoming negative for hexadecane.  相似文献   

3.
A new chelate (η5-C5H5)2Ti(SB)2, whereSB=O, N donor Schiff base salicylidene-4-methylaniline, was synthesized. The course of thermal degradation of the chelate was studied by thermogravimetric (TG) and differential thermal analysis (DTA) under dynamic conditions of temperature. The order of the thermal decomposition reaction and energy of activation was calculated from TG curve while from DTA curve the change in enthalpy was calculated. Evaluation of the kinetic parameters was performed by Coats-Redfern as well as Piloyan-Novikova methods which gaven=1, ΔH=1.114 kJ·mol?1, ΔE=27.01 kJ·mol?1, ΔS=?340.12 kJ·mol?1·K?1 andn=1, ΔH=1.114 kJ·mol?1, ΔE=20.01 kJ·mol?1, ΔS=?342.60 kJ·mol?1·K?1, respectively. The chelate was also characterized on the basis of different spectral studies viz. conductance, molecular weight, IR, UV-visible and1H NMR, which enabled to propose an octahedral structure to the chelate.  相似文献   

4.
The effect of In impurity on the crystallization kinetics and the changes taking place in the structure of (Se7Te3) have been studied by DTA measurements at different heating rates (α=5 deg·min?1, 10 deg·min?1, 15 deg·min?1 and 20 deg·min?1). From the heating rate dependence of the values ofT g,T c andT p, the glass transition activation energy (E t) and the crystallization activation energy (E c) have been obtained for different compositions of (Se7Te3)100?xInx (0≤×≤20). The variation of viscosity as a function of temperature has been evaluated using Vogel-Tamman-Fulcher equation. The crystallization data are analysed using Kissinger's and Matusita's approach for nonisothermic crystallization. It has been found that for samples containing In=0, 10, 15, 20 at%, three dimensional nucleation is predominant whereas for samples containing In=5 at%, two dimensional nucleation is the dominant mechanism. The compositional dependence ofT g and crystallization kinetics are discussed in terms of the modification of the structure of the Se?Te system.  相似文献   

5.
The product from reaction of samarium chloride hexahydrate with salicylic acid and Thioproline, [Sm(C7H5O3)2·(C4H6NO2S)]·2H2O, was synthesized and characterized by IR, elemental analysis, molar conductance, and thermogravimetric analysis. The standard molar enthalpies of solution of [SmCl3·6H2O(s)], [2C7H6O3(s)], [C4H7NO2S(s)] and [Sm(C7H5O3)2·(C4H7NO2S)·H2O(s)] in a mixed solvent of absolute ethyl alcohol, dimethyl sulfoxide(DMSO) and 3 mol L?1 HCl were determined by calorimetry to be Δs H m Φ [SmCl3 δ6H2O (s), 298.15 K]= ?46.68±0.15 kJ mol?1 Δs H m Φ [2C7H6O3 (s), 298.15 K]= 25.19±0.02 kJ mol?1, Δs H m Φ [C4H7NO2S (s), 298.15 K]=16.20±0.17 kJ mol?1 and Δs H m Φ [Sm(C7H5O3)2·(C4H6NO2S)]·2H2O (s), 298.15 K]= ?81.24±0.67 kJ mol?1. The enthalpy change of the reaction (1) $$ SmCl_3 \cdot 6H_2 O(s) + 2C_7 H_6 O_3 (s) + C_4 H_7 NO_2 S(s) = Sm(C_7 H_5 O_3 )_2 \cdot (C_4 H_6 NO_2 S) \cdot 2H_2 O(s) + 3HCl(g) + 4H_2 O(1) $$ was determined to be Δs H m Φ =123.45±0.71 kJ mol?1. From date in the literature, through Hess’ law, the standard molar enthalpy of formation of Sm(C7H5O3)2(C4H6NO2S)δ2H2O(s) was estimated to be Δs H m Φ [Sm(C7H5O3)2·(C4H6NO2S)]·2H2O(s), 298.15 K]= ?2912.03±3.10 kJ mol?1.  相似文献   

6.
The molecular structure and conformation of 2,3-dichloro-1-propene have been determined by gas-phase electron diffraction at nozzle temperatures of 24, 90 and 273°C. The molecules exist as a mixture of two conformers with the chlorine atoms anti (torsion angle ∠φ = 0°) or gauche (∠φ = 109°) to each other and with the anti form the more stable. The composition (mole fraction) of the vapor with uncertainties estimated at 2σ was found to be 0.55 (0.08), 0.49 (0.08) and 0.41 (0.10) at 24, 90 and 273°, respectively. These values correspond to an energy difference with estimated standard deviation ΔE° = E°g-E°a = 0.7 ± 0.3 kcal mol?1 and an entropy difference ΔS° = S°g-S°a = 0.6 ± 0.9 cal mol?1 K?1. Some of the diffraction results, together with spectroscopic observations, permit the evaluation of an approximate torsional potential function of the form 2V = V1 (1 - cos φ) + V2 (1 - cos 2φ) + V3 (1 - cos 3φ); the results are V1 = 4.4 ± 0.5, V2 = ?2.9 ± 0.5 and V3 = 4.8 ± 0.2, all in kcal mol?1. The results at 24°C for the distance (ra) and angle (∠α) parameters, with estimated uncertainties of 2σ, are: r(Csp2-H) = 1.098(0.020)Å, r(Csp3-H) = 1.103(0.020)Å, r(CC) = 1.334(0.009)Å, r(C-C) = 1.504(0.013)Å, r(Csp2-Cl) = 1.752(0.021)Å, r(Csp3-Cl) = 1.776(0.020)Å, ∠C-CC = 127.6(1.1)°, ∠Csp3-Csp2-Cl = 110.2(1.0), ∠Csp2-Csp3-Cl = 113.1(1.2)°, ∠H-Csp3-H = 109.5° (assumed), ∠CC-H = 120.0° (assumed) and ∠φ = 108.9(3.4)°.  相似文献   

7.
It has been confirmed by 1H and 13C NMR spectroscopies that Sn(σ-C7H7)Ph3 undergoes either 1,4- or 1,5-shifts of the SnPh3 moiety around the cycloheptatrienyl ring with ΔH3 = 13.8 ± 0.4 kcal mol?1, ΔS3 = ?5.6 ± 1.2 cal mol?1 deg?1, and ΔG3300 = 15.44 ± 0.14 kcal mol?1. Similarly, (σ-5-cyclohepta-1,3-dienyl)triphenyltin undergoes 1,5-shifts with ΔH3 = 12.4 ± 0.6 kcal mol?1, ΔS3 = ?11.2 ± 1.8 cal mol?1 deg?1, and ΔG3300 = 15.76 ± 0.13 kcal mol?1. It is therefore probable that Sn(σ-5-C5H5)R3 and Sn(σ-3-indenyl)R3 do not undergo 1,2-shifts as previously suggested but really undergo 1,5-shifts.  相似文献   

8.
The formation of complexes at pH 4.7 of the Hg(II) with five monothiosemicarbazone and two dithiosemicarbazone has been studied. The mercury(II) reacts with monothiosemicarbazones of salicylaldehyde (λmax = 363 nm, E = 1.69 × 104liters · mol?1cm?1), pi-colinadehyde (λmax = 363 nm, E = 2.38 × 104liters · mol?1cm?1), 6-methyl-picolinaldehyde (λmax = 363 nm, E = 2.28 × 104liters · mol?1cm?1), di-2-pyridylketone (λmax = 380 nm, E = 2.08 × 104liters · mol?1cm?1), and o-naphthoquinone (λmax = 540 nm, E = 1.03 × 104liters · mol?1cm?1) and with dithiosemicarbazones of 1,4-dihydroxyphthalimide (λmax = 430 nm, E = 2.56 × 104liters · mol?1cm?1) and dipyridylglyoxal (λmax = 363 nm, E = 2.37 × 104liters · mol?1cm?1). A critical comparison of the stoichiometry and apparent stability constant of complexes with mono- and dithiosemicarbazones is given.  相似文献   

9.
The low-temperature (5 to 310 K) heat capacity of cesium fluoroxysulfate, CsSO4F, has been measured by adiabatic calorimetry. At T = 298.15 K, the heat capacity Cpo(T) and standard entropy So(T) are (163.46±0.82) and (201.89±1.01) J · K?1 · mol?1, respectively. Based on an earlier measurement of the standard enthalpy of formation ΔHfo the Gibbs energy of formation ΔGfo(CsSO4F, c, 298.15 K) is calculated to be ?(877.6±1.6) kJ · mol?1. For the half-reaction: SO4F?(aq)+2H+(aq)+2e? = HSO4?(aq)+HF(aq), the standard electrode potential E at 298.15 K, is (2.47±0.01) V.  相似文献   

10.
Vanadium(II) ions form with the pyridine-2-carboxylate ligand a deep blue, tris-substituted complex absorbing at 660 nm (ε = 7.2 × 103 M?1) cm?1) with a shoulder at 450 nm. Reversible spectroelectrochemistry and cyclic voltammetry were observed for this complex, with E12 = ?0.448 V vs NHE, and ΔSrcθ = ?6 cal · mol?1 · deg?1. Electron transfer kinetics with [CO(en)3]3+ led to k12 = 3100 M?1 s?, ΔH = 12.4 kcal · mol?1 and ΔS = ?0.9 cal · mol?1 · deg?1 (I = 0.10 M). For the related [Co(NH3)6]3+ complex, k13 = 1.9 × 104 M?1 s?1. The self-exchange rate constant and activation parameters were analysed in terms of relative Marcus theory.  相似文献   

11.
The product from reaction of lanthanum chloride heptahydrate with salicylic acid and thioproline, [La(Hsal)2•(tch)]•2H2O, was synthesized and characterized by IR, elemental analysis, molar conductance, thermogravimatric analysis and chemistry analysis. The standard molar enthalpies of solution of LaCl3•7H2O (s), [2C7H6O3 (s)], C4H7NO2S (s) and [La(Hsal)2•(tch)]•2H2O (s) in a mixed solvent of absolute ethyl alcohol, dimethyl sulfoxide (DMSO) and 3 mol•L-1 HCl were determined by calorimetry to be [LaCl3•7H2O (s), 298.15 K]=(-102.36±0.66) kJ•mol-1, [2C7H6O3 (s), 298.15 K]=(26.65±0.22) kJ•mol-1, [C4H7NO2S (s), 298.15 K]=(-21.79±0.35) kJ•mol-1 and {[La(Hsal)2•(tch)]•2H2O (s), 298.15 K}=(-41.10±0.32) kJ•mol-1. The enthalpy change of the reaction LaCl3•7H2O (s)+2C7H6O3 (s)+C4H7NO2S (s)=[La(Hsal)2•(tch)]•2H2O (s)+3HCl (g)+5H2O (l) (Eq. 1) was determined to be =(41.02±0.85) kJ•mol-1. From date in the literature, through Hess’ law, the standard molar enthalpy of formation of [La(Hsal)2•(tch)]•2H2O (s) was estimated to be {[La(Hsal)2•(tch)]•2H2O (s), 298.15 K}=(-3017.0±3.7) kJ•mol-1.  相似文献   

12.
The dissociation energy of the C2H4 · HCl van der Waals complex was determined to be 3.18±0.73 kcal mol?1 by a dissociative photoionization technique. C2H4 · HCl was produced by free expansion of a 1:4 mixture of C2H4 in HCl and the clusters were ionized with tunable synchrotron radiation. The photoionization efficiency function of (C2H4 · HCl)+ from C2H4 · HCl was determined between 600 and 1,300 Å and the onset for (C2H4 · HCl)+ was established as 1,163±2 Å = 10.66±0.02 eV; these values give ΔH f 0 (C2H4 · HCl) = ?10.7±0.7 kcal mol?1 and ΔH f 0 (C2H4·HCl+)=235.1±0.9 kcal mol?1. A complex ion dissociation energyD 0(C2H4 · HCl+) = ?0.3±0.9 kcal mol?1 was calculated from the results. The major features on the PIE curve for C2H4 · HCl+ can be analyzed in terms of the known energetic features of C2H 4 + and HCl. An extended energy diagram for the C2H4 + HCl system is presented.  相似文献   

13.
Heat capacities C p(T) of L-valine and DL-valine were measured in the temperature range 6–300 K with an adiabatic calorimeter; thermodynamic functions were calculated based on these measurements. At 298.15 K, the values of heat capacity, C p; entropy, S m 0 (T) ? S m 0 (0); enthalpy, H m 0 (T) ? H m 0 (0) of L-valine are equal, respectively, to 167.9 ± 0.3 J K?1 mol?1; 178.5 ± 0.4 J K?1 mol?1; and 27510 ± 60 J mol?1. For DL-valine, these values are equal, respectively, to 167.3 ± 0.3 J K?1 mol?1, 174.4 ± 0.3 J K?1 mol?1, and 27000 ± 50 J mol?1. The difference between the heat capacities of enantiomer and racemate has been calculated and compared with the similar data for serines, cysteines, and phenylglycines.  相似文献   

14.
The enthalpy change of the reaction at 298 K between Br2 (l) and Sn(c) in CS2 as solvent giving SnBr4 (s) has been determined by calorimetry to be (?374, 2±1.4) kJ·mol?1, [(?89.45±0.33) kcal·mol?1]. By the same method the heat of solution of SnBr4 (c) in CS2 has been found to be (11.9±0.3) kJ·mol?1, [(2.84±0.08) kcal·mol?1]. Combining these results, a value of (?386.1±1.5) kJ·mol?1, [(?92.3±0.4) kcal·mol?1] is derived for the standard heat of formation of SnBr4 (c). Substituting this figure in the thermochemical cycle hitherto used for calculating the heat of formation of SnBr4 (c) gives ?124.3 kcal·mol?1 for the standard heat of formation of SnCl4 (l), which is in reasonable agreement with a recent determination of this quantity8.  相似文献   

15.
16.
For n-decane (p, Vm, T) has been determined from 298 to 673 K and from atmospheric pressure to 300 MPa by measuring p(T) at constant molar volumes Vm from 190 to 1050 cm3·mol?1. The experimental results are tabulated and represented by equations. Thermal pressure coefficients have been calculated.  相似文献   

17.
Reactions of Fe+ and FeL+ [L=O, C4H6, c-C5H6, C5H5, C6H6, C5H4(=CH2)] with thiophene, furan, and pyrrole in the gas phase by using Fourier transform mass spectrometry are described. Fe+, Fe(C5H5)+, and FeC6H 6 + yield exclusive rapid adduct formation with thiophene, furan, and pyrrole. In addition, the iron-diene complexes [FeC4H 6 + and Fe(c-C5H6)+], as well as FeC5H4(=CH2)+ and FeO+, are quite reactive. The most intriguing reaction is the predominant direct extrusion of CO from furan by FeC4H6 +, Fe(c-C5H6)+, and FeC5H4(=CH2)+. In addition, FeC4H 6 + and Fe(c-C5H6)+ cause minor amounts of HCN extrusion from pyrrole. Mechanisms are presented for these CO and HCN extrusion reactions. The absence of CS elimination from thiophene may be due to the higher energy requirements than those for CO extrusion from furan or HCN extrusion from pyrrole. The dominant reaction channel for reaction of Fe(c-C5H6)+ with pyrrole and thiophene is hydrogen-atom displacement, which implies DO(Fa(N5H5)+-C4H4X)>DO(Fe(C5H5)+-H)=46±5 kcal mol?1. DO(Fe+-C4H4S) and DO(Fe+-C4H5N)=DO(Fe+-C4H6)=48±5 kcal mol?1. Finally, 55±5 kcal mol?1=DO(Fe+-C6H6)>DO(Fe+-C4H4O)>DO(Fe+-C2H4)=39.9±1.4 kcal mol?1. FeO+ reacts rapidly with thiophene, furan, and pyrrole to yield initial loss of CO followed by additional neutral losses. DO(Fe+-CS)>DO(Fe+-C4H4S)≈48±5 kcal mol?1 and DO(Fe+-C4H5N)≈48±5 kcal mol?1>DO(Fe+-HCN)>DO(Fe+-C2H4)=39.9±1.4 kcal mil?1.  相似文献   

18.
Treatment of [BzPh3P][AuCl2] with [Hg(x-C6H4NO2)2] (x = o, m, or p) gives anionic gold(I) complexes of the type [BzPh3P][Au(R)Cl](R = o-, m- or p-C6H4NO2, Bz = C6H5CH2). The chloro ligand in [Au(o-C6H4NO2)Cl]? can be replaced by bromo or iodo ligands by use of NaBr or NaI. The anions [Au(R)Cl]? react with neutral monodentate ligands, L, to give neutral mononuclear complexes [Au(R)L] (R = o-C6H4NO2, L = PPh3, AsPh3; R = m-C6H4NO2, L = PPh3) and with 1,2-bis(diphenylphosphino)ethane (dpe) to give [Au2(R)2(dpe)] (R = o-C6H4NO2). The corresponding [Au(p-C6H4NO2)Cl]? reacts with PPh3 or AsPh3 to give mixtures containing [AuClL]. The anionic ortho-nitrophenylgold(I) complex is much more stable than its meta- or para-nitrophenyl isomers. These are thought to be the first reports of nitrophenylgold(I) complexes.  相似文献   

19.
Final products of isothermal pyrolysis of CF2HBr, CF2ClBr, CH3I, CH2I2, CHI3, i-C3F7I, and C4F9I were determined and mechanisms of their formation were proposed. The enthalpy of formation of the free biradical CFBr··fH 0 0 = 80±20, Δ fH 0 298 = 60±20 kJ mol?1) was estimated. The dissociation energies ED, 0(ID2C-D), ED,298(ID2C-D), and ED,0(IH2C-D) equal to 437±6, 444±6, and 435±4 kJ mol?1, respectively, were determined.  相似文献   

20.
The reaction of alkyl aryl N-p-tosylsulphilimines with thiophenolate ion was found to afford quantitatively the sulphide that arises by an SN2 like reaction on the carbon atom adjacent to the tri-valent sulphur atom. This reaction was also found to proceed smoothly with such compounds as sulphoxides and sulphones and sulphoxmanes. The kinetic study on the reaction between aryl methyl N-p-tosylsulphilimine with thiophenolate ion in DMF reveals that the reaction is of second order, namely, first order with respect to each thiophenolate ion and the sulphilimine. The enthalpy and entropy of activation for the reaction are ΔH = ?17· kcal/mol and ΔS = ?5·7 eu respectively. The effect of substituents in the reaction, p-XC6H4+(?SO2C6H4Y-p)CH3 + p-ZC6H4SK is nicely correl with Hammett σ values giving ?x = + 2·4, ?y = + 1·2 and ?z = ?1·8 respectively. Meanwhile, a marked steric retardation by a bulky alkyl group in alkyl phenyl N-p-tosylsulphilimine is observed. Furthermore, from the stereochemical study of the reaction using an optically active sec-octyl phenyl N-p-tosylsulphilimine with thiophenolate ion it is concluded that the reaction proceeds via a typical SN2 process on α-carbon atom attached to the tri-valent sulphur atom.  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