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 984 毫秒
1.
The following copper(I) and silver(I) complexes of 2-amino-1,3,4-thiadiazole (atz) and 2-ethylamino-1,3,4-thiadiazole (eatz) have been prepared and studied by conductometric, IR and Raman methods: CuXL(X = Cl, Br, I; L = atz, eatz), CuXL3(X = ClO4, NO3; L = atz, eatz), AgClO4·1.5atz·1/3 EtOH, AgNO3·2.5atz, AgClO4·3eatz, AgNO3·eatz. The ligands are bonded through the amine nitrogen atoms with ν(MN) bands in the 520–410 cm?1 region. The CuXL complexes have a trigonal (N, 2Xb) coordination with a probable weaker axial interaction. The CuXL3 and AgCIO4·3eatz complexes probably have a trigonal pyramidal (3N,O) coordination. In the atz complexes of silver perchlorate and nitrate some ligand molecules are bridging. The AgNO3·2.5atz complex is likely to have a dimeric structure with tetrahedral coordination of the silver ion.  相似文献   

2.
The densities and volumetric specific heats of aqueous solutions of Bu3NHBr, Pent3NHCl, and three diazonium salts, HN?Oct3?NHBr2, HN?Dec3?NHCl2, and Bu3N?Oct?NBu3Br2, have been measured at 25°C. From these data, the apparent molal volumes φ v and apparent molal heat capacities φ c have been calculated and are reported here. In the series of compounds chosen, the diazonium (higher homologs) can be regarded as dimers of the alkyl-substituted ammonium ions (lower homologs), and these systems are examined as chemical models for the hydrophobic interaction. With the three homologous pairs studied here, the chemical model predicts that the strong interaction (limitingly, chemical binding) of two hydrocarbon chains in water leads to a major decrease in both φ v and φ c of the interacting solutes, ca.?22 cm3-mole?1 and ?200 J-oK?1-mole?1. These predictions constitute limiting behavior — useful, but not sufficient, to explain the observed concentration dependence of φ v and φ c in aqueous solutions of the lower homologs Bu3NHBr, Pent3NHCl, and Bu4NBr. An explanation for the concentration dependence of φ c is suggested with reference to ultrasonic relaxation data.  相似文献   

3.
The heat capacity of a sample of Cs2CrO4 was determined in the temperature range 5 to 350 K by aneroid adiabatic calorimetry. The heat capacity at constant pressure Cpo(298.15 K), the entropy So(298.15 K), the enthalpy {Ho(298.15 K) - Ho(0)} and the function ? {Go(298.15 K) - Ho(0)}298.15K were found to be (146.06 ± 0.15) J K?1 mol?1, (228.59 ± 0.23) J K?1 mol?1, (30161 ± 30) J mol?1, and (127.43 ± 0.13) J K?1 mol?1, respectively. The heat capacity Cpo(298.15 K) and entropy So(298.15 K) and entropy So(298.15 K) of Rb2CrO4 are estimated to be (146.0 ± 1.0) J K?1 mol?1 and (217.6 ± 3.0) J K?1 mol?1, respectively.  相似文献   

4.
The standard enthalpy of combustion of cyclohexylamine has been measured in an aneroid rotating-bomb calorimeter. The value ΔHoo(c-C6H11NH2, 1) = ?(4071.3 ± 1.3) kJ mol?1 yields the standard enthalpy of formation ΔHfo(c-C6H11NH2, 1) = ?(147.7 ± 1.3) kJ mol?1. The corresponding gas-phase standard enthalpy of formation for cyclohexylamine is ΔHfo(c-C6H11NH2, g) = ?(104.9 ± 1.3) kJ mol?1. The standard enthalpy of formation of cyclohexylamine hydrochloride, ΔHfo(c-C6H11NH2·HCl, c) = ?(408.2 ± 1.5) kJ mol?1, was derived by combining the measured enthalpy of solution of the salt in water, literature data, and the ΔHco measured in this study. Comment is made on the thermochemical bond enthalpy H(CN).  相似文献   

5.
A thoroughly analyzed specimen of β-uranium disulfide of composition US1.992±0.002 has been studied by fluorine-bomb calorimetry. The standard molar energy of combustion: ΔcUmo(US1.992, cr, β, 298.15 K) = ?(4092.5±7.5) kJ·mol?1 has been determined on the basis of the reaction: US1.992(cr, β) + 8.976F2(g) = UF6(cr) + 1.992F6(g). The standard molar enthalpy of formation: ΔfHmo(US1.992, cr, β, 298.15 K) = ?(519.7±8.0) kJ·mol?1 was derived, and from that result ΔfHmo(US2, cr, 298.15 K) = ?(521±8) kJ·mol?1 is estimated.  相似文献   

6.
Chloroacetyl chloride is studied by gas-phase electron diffraction at nozzle-tip tempera- tures of 18, 110 and 215°C. The molecules exist as a mixture of anti and gauche confor- mers with the anti form the more stable. The composition (mole fraction) of the vapor with uncertainties estimated at 2σ is found to be 0.770 (0.070), 0.673 (0.086) and 0.572 (0.086) at 18, 110 and 215°C, respectively. These values correspond to an energy difference with estimated standard deviation ΔEo = Eog -Eoa = 1.3 ± 0.4 kcal mol?1 and an entropy difference ΔSo = Sog -Soa = 0.7 ± 1.1 cal mol?1 K?1. Certain of the diffraction results 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 = 1.19 ± 0.33, V2 = 0.56 ± 0.20 and V3 = 0.94 ± 0.12, all in kcal mol?1. The results for the distance (ra), angle (∠α) and r.m.s. amplitude parameters obtained at the three temperatures are entirely consistent. At 18°C the more important parameters are, with estimated uncertainties of 2σ, r(C-H) = 1.062(0.030) Å, r(CO) = 1.182(0.004) Å, r(C-C) = 1.521(0.009) Å. r(CO-Cl) = 1.772(0.016) Å, r(CH2-Cl) = 1.782(0.018) Å, ∠C-C-0 = 126.9(0.9)°, ∠CH2-CO-C1 = 110.0(0.7)°,∠CO-CH2-C1 = 112.9(1–7)°, ∠H-C-H = 109.5° (assumed), ∠φ (gauche torsion angle relative to 0° for the anti form) = 116.4(7.7)°, δ (r.m.s. amplitude of torsional vibration in the anti conformer) == 17.5(4.2)°.  相似文献   

7.
An accurate gas-phase acidity for germane (enthalpy scale, equivalent to the proton affinity of GeH3 ?), ΔH acid o(GeH4) = 1502.0 ± 5.1 kJ mol?1, is obtained by constructing a consistent acidity ladder between GeH4, and H2S by using Fourier transform-ion cyclotron resonance spectrometry, and 0 and 298.15 K values for the first bond dissociation energy of GeH4 are proposed: D0 o(H3Ge-H) = 352 ± 9 kJ mol?1; D o(H3Ge-H) = 358 ± 9 kJ mol?1, respectively. These results are compared with experimental and theoretical data reported in the literature. Methylgermane was found to be a weaker acid than germane by approximately 35 kJ mol?1: ΔH acid o = 1536.6 kJ mol?1.  相似文献   

8.
Diffraction data on BaI2, analyzed by a new approach, indicate an anharmonic potential with a barrier of 71(12) cm?1 at a linear geometry. The structural and vibrational parameters were found to be reh(Ba-Io) = 3.150(7)Å, ∠eIBaI = 148.0(9) °, fq = 0.69(8) mdyn/Å,fqq= 0.14(6) mdyn/Å, k2 = ?0.0075(15) mdyn/Å, k4 = 0.0025(9) mdyn/Å3, v1 = 106(12) cm?1 and v3 = 145(21) cm?1. The bending frequency v2 is predicted to be near 16 cm?1.  相似文献   

9.
Enthalpies of the synthesis reactions of the two compounds KCdCl3(cr) and K4CdCl6(cr) from KCl(cr) and CdCl2(cr) have been measured by drop calorimetry of solid samples of KCl, CdCl2, KCdCl3, and K4CdCl6 into melted mixtures of KCl and CdCl2. For the two reactions: (1), CdCl2(cr)+KCl(cr) = KCdCl3(cr); and (2), CdCl2(cr)+4KCl(cr) = K4CdCl6(cr), the experiments lead to the two standard molar enthalpies of reaction at 298.15 K: Δ1Hmo = ?(21.7±1.0) kJ·mol?1 and Δ2Hmo = ?(39.0±3.8) kJ·mol?1. These values are not in good agreement with those of previous workers.  相似文献   

10.
The silver(I) nitrate complexes with 2,3-, 2,4-, 2,6-, and 3,5-lutidine (Lut, dimethylpyridine C7H9N), [AgNO3(Lut)2], are synthesized and studied by multinuclear NMR (1H, 13C, and 15N) in various solvents (chloroform, dimethyl sulfoxide, and acetonitrile). The influence of steric and electronic factors of the organic ligand on the parameters of the NMR spectra is revealed. It is shown that the 15N NMR spectra are the most informative. The structure of complex [AgNO3(3,5-Lut)2] is determined. The crystals are monoclinic, space group C2/c, a = 14.599(1) Å, b = 8.422(1) Å, c = 12.954(1) Å, β = 99.60(1)°, V = 1570(2) Å3, ρcalcd = 1.625 g/cm3, Z = 4. The structure is built of discrete neutral complexes [AgNO3(3,5-Lut)2]. The coordination mode of the Ag+ ion includes two nitrogen atoms of two crystallographically equivalent lutidine ligands (Ag-N 2.194(5) Å, angle NAgN 147.6(3)°). The nitrate ion behaves as a weak chelating ligand with respect to the Ag+ ion (Ag…O 2.674(6) Å).  相似文献   

11.
Equilibrium vapor pressures were determined at temperatures between 294 K and 353 K for the NiCl2-CH3CN system.Compositions studied ranged from molar ratios of CH3CN to NiCl2 of 0.27 to 1.90. Three stoichiometric compounds were identified: NiCl2(CH3CN)2, NiCl2(CH3CN), NiCl2(CH3CN)0.88. Per mole of gaseous CH3CN the values of ΔHo and ΔSo were calculated to be 52.0 ±0.4 kJ mol?1 and 149.0±1.3 J mol?1 K?1 for the decomposition of NiCl2(CH3CN)2, and 25.9 ±0.8 kJ mol?1 and 58.6± 2.1 J mol?1 K?1 for the decomposition of NiCl2(CH3CN). Below a composition of NiCl2(CH3CN)0.88 the phase diagram is complex and could not be interpreted in terms of specific stoichiometric compounds.  相似文献   

12.
The apparent molal volume φv, expansibility φE, compressibility φK, and heat capacity φc of NaCl were measured in urea-water mixtures, as a function of salt (<1.5m) and urea (<13m) concentrations at 25°C. At a fixed urea concentration, the transfer functions from H2O to 3m urea are linear functions of the NaCl aquamolality. At a fixed salt aquamolality, (0.1m), the sign of the transfer functions is in the direction of a decrease in the structure-breaking effect, and the absolute values of the transfer functions tend to level off at high urea concentrations (13m). The functions φv, φE, φK, φc, and (?φv/?T)p were measured for the sodium halides and alkali, bromides (chlorides in the case of φK) at a fixed salt aquamolality of 0.1m and fixed urea molality of 3m. The corresponding transfer functions from H2O to 3m urea are opposite those from H2O to D2O and similarly are relatively independent of ionic size. This suggests that urea, shows no specific interaction affinity for ions and that the overall number of water molecules influenced by the ions is relatively constant for all alkali halides. The lithium halides are an exception in that Li+ seems to have hardly any structure-breaking effect.  相似文献   

13.
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   

14.
Isomerically pure endohedral metallofullerene Dy@C82(C 2v) was synthesized by the electric arc method, extracted from the soot with o-dichlorobenzene, isolated from the extract by HPLC, and characterized by mass spectrometry and spectrophotometry. The spectrophotometric titration of a solution of endohedral metallofullerene Dy@C82(C 2v) was conducted with potassium perchlorotriphenylmethide. The concentration of Dy@C82(C 2v) in o-dichlorobenzene was determined, and the molar absorption coefficients for its neutral and anionic forms were calculated (3.0?103 (at 927 nm) and 4.0?103 mol–1 L cm–1 (at 884 nm), respectively.  相似文献   

15.
From measurements of the heats of iodination of CH3Mn(CO)5 and CH3Re(CO)5 at elevated temperatures using the ‘drop’ microcalorimeter method, values were determined for the standard enthalpies of formation at 25° of the crystalline compounds: ΔHof[CH3Mn(CO)5, c] = ?189.0 ± 2 kcal mol?1 (?790.8 ± 8 kJ mol?1), ΔHof[Ch3Re(CO)5,c] = ?198.0 ± kcal mol?1 (?828.4 ± 8 kJ mo?1). In conjunction with available enthalpies of sublimation, and with literature values for the dissociation energies of MnMn and ReRe bonds in Mn2(CO)10 and Re2(CO)10, values are derived for the dissociation energies: D(CH3Mn(CO)5) = 27.9 ± 2.3 or 30.9 ± 2.3 kcal mol?1 and D(CH3Re(CO)5) = 53.2 ± 2.5 kcal mol?1. In general, irrespective of the value accepted for D(MM) in M2(CO)10, the present results require that, D(CH3Mn) = 12D(MnMn) + 18.5 kcal mol?1 and D(CH3Re) = 12D(ReRe) + 30.8 kcal mol?1.  相似文献   

16.
The heat capacities of potassium, rubidium, cesium, and thallium azides were determined from 5 to 350 K by adiabatic calorimetry. Although the alkali-metal azides studied in this work exhibited no thermal anomalies over the temperature range studied, thallium azide has a bifurcated anomaly with two maxima at (233.0±0.1) K and (242.04±0.02) K. The associated excess entropy was 0.90 calth K?1 mol?1. The thermal properties of the azides and the corresponding structurally similar hydrogen difluorides are nearly identical. Both have linear symmetrical anions. However, thallium azide shows a solid-solid phase transition not exhibited by thallium hydrogen difluoride. At 298.15 K the values of Cpo, So, and ?{Go(T)?Ho(0)}T, respectively, are 18.38, 24.86, and 12.676 calth K?1 mol?1 for potassium azide; 19.09, 28.78, and 15.58 calth K?1 mol?1 for rubidium azide; 19.89, 32.11, and 18.17 calth K?1 mol?1 for cesium azide; and 19.26, 32.09, and 18.69 calth K?1 mol?1 for thallium azide. Heat capacities at constant volume for KN3 were deduced from infrared and Raman data.  相似文献   

17.
The vapour pressure and orthobaric molar volumes of tetramethylsilane have been messured from 373 K to the critical temperature, and the critical temperature (IPTS-68: To = 448.64 K), critical pressure (pc = 2821 kPa), and critical molar volume (Vmc = 361 cm3 mol?1) have been determined.  相似文献   

18.
There are calculated for 298 K and higher temperatures the standard entropies of the liquid and the important gas species (monomer, dimer, and trimer) of MoF5, and their differences in enthalpy and Gibbs energy. Besides estimated mean heat capacities, the input data are: the vapor pressure at one temperature and the saturated-vapor density at two temperatures (based on measurements in this laboratory); the standard entropies of the three gas species (estimated from published spectroscopic data on the crystal and monomer); and an ion-abundance proportion (from a published mass-spectral study). This ion proportion is corrected for fragmentation after demonstrating that changes in the ion proportions for a similar fluoride, RhF5, can be explained by assuming that the rates of fragmentation of the dimer and trimer are equal. For those input data considered most uncertain (the vapor pressure, the entropies of dimer and trimer, and the ratios of mass-spectral ionization efficiencies) several sets of values covering their respective estimated ranges of uncertainty are used alternatively in the calculations. Results at 298.15 K—expressed in the form “preferred value (probable range due to uncertainty)”, with Ho and Go relative to the liquid, and So—are as follows: liquid, So = 46 (44.5 to 49) calth K?1 mol?1; monomer, Ho = 19 (17 to 22) kcalth mol?1, Go = 8.5 (7.5 to 11.5) kcalth mol?1, So = 80.6 calthK?1; dimer, Ho = 15.7 (15 to 16) kcalth mol?1, Go = 6.48 (6.45 to 6.51) kcalth mol?1, So = 123 (121 to 127) calth K?1 mol?1; trimer, Ho = 17.7 (16 to 21) kcalth mol?1, Go = 8.4 (8.1 to 8.7) kcalth mol?1, So = 169 (161 to 185) calth K?1 mol?1. The calculated mole fractions of monomer, dimer, and trimer in the saturated vapor at 298.15 K average 0.027, 0.94, and 0.035, respectively.  相似文献   

19.
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   

20.
The heat capacities of LaCl3, PrCl3, and NdCl3 have been measured from 5 to 350 K by adiabatic calorimetry. No co-operative thermal anomalies were seen in the temperature range investigated but substantial magnetic heat-capacity contributions of the non-cooperative (Schottky) type were found. Subtraction of the heat capacity of the diamagnetic and isostructural LaCl3 from those of the paramagnetic members yields experimental Schottky heat-capacity contributions which are compared with heat capacities derived from spectroscopically determined energy levels. Small discrepancies between the calculated and experimental contributions are probably due to differences in lattice heat capacities between LaCl3 and the others. The values of {(So(298.15 K) ? So(0)}/calth K?1 mol?1 are for LaCl3 and NdCl3, 32.88 and 36.67. Due to the possibility of a low-temperature phase transition, the entropy of PrCl3 covers only the experimental range of this research and that of Colwell, Mangum, and Utton. {So(298.15 K) ? So(0.294 K)} for PrCl3 is 36.64 calth K?1 mol?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