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 140 毫秒
1.
Reaction of UV and UVI Compounds with SOCl2 UO3, UO2Cl2, UCl6, and UCl5 reacted with OSCl2 yield always UCl5 · SCl2, [SCl3]+ [UCl6]? or a mixture of these compounds, but not an adduct UCl5 · OSCl2. An X-ray study was carried out with single crystals of [SCl3]+[UCl6]?. It crystallizes in the orthorhombic space group P212121 with the lattice constants a = 1066.8, b = 1071.2, c = 1133.3 pm and with Z = 4, containing isolated pyramidal SCl3+ (rSCl = 196.2 ± 1.1 pm ?SCl2 = 102.34 ± 1.13°) and octahedral UCl6? ions (rUCl = 251.1 ± 2.6 pm).  相似文献   

2.
《Tetrahedron》1988,44(2):405-416
The synthesis, crystal structure (7,8), conformation and dynamics of pentaspiro[2.0.2.0.2.0.2.0.2.1] hexadecane 6, pentaspiro [3.0.2.0.3.0.2.0.3.1] nonadecane 7 and pentaspiro [3.0.3.0.3.0.3.0.3.1] heneicosane 8 are described. Chair conformations have been found in the solid state (7,8) and in solution (6,7,8). The activation parameters of the chair-to-chair interconversion have been determined from bandshape analyses of exchange broadened 1H-NMR (6,7) and13 C-NMR spectra (8), respectively. The results were as follows: 6: ΔH3 = 48.9 ± 0.8 kJ/mol, ΔS3 = -20.7 ± 2.8 J/mol, grd, ΔG3298 = 55.0 ± 0.1 kJ/mol; 7: ΔH3=51.2±0.7 kJ/mol, ΔS3 = -12.0±2.4 J/mol, grd, ΔG3298 = 54.8±0.1 kJ/mol; 8: ΔH3 - 74.2±0.6 kJ/mol, ΔS3 =-21.9 ± 1.5 J/mol, grd, ΔG3298 = 80.7 ± 0.2 kJ/mol. On the basis of these values the barrier of inversion of the still unknown hexaspirane 5 is predicted to exceed 160 kJ/mol.  相似文献   

3.
The Tl-Te-Cl system was studied in the Tl-TlCl-Te composition region by differential thermal analysis, X-ray powder diffraction, and emf and microhardness measurements. A series of polythermal sections, an isothermal section at 400 K, and a projection of the liquidus surface of the phase diagram were constructed. The ternary compound Tl5Te2Cl characterized by a wide homogeneity region and incongruent melting by a syntectic reaction at 708 K was shown to exist. This compound was found to crystallize in tetragonal lattice (space group I4/mcm) with the parameters a = 8.921 Å, c = 12.692 Å, Z = 4. Wide phase separation regions were also found in the system, including a three-phase separation region in the Tl-TlCl-Tl2Te subsystem. Regions of primary crystallization of phases, and the types and coordinates of in- and monovariant equilibria in the T-x-y diagram were determined. From emf measurement data, the standard thermodynamic functions of formation and the standard entropy were calculated for the compound Tl5Te2Cl, as follows: ?ΔG 298 0 = 355.9 ± 1.1 kJ/mol, ?ΔH 298 0 = 377.1 ± 5.0 kJ/mol, and S 298 0 = 474.1 ± 6.8 J/(mol K).  相似文献   

4.
《Thermochimica Acta》1987,114(2):303-311
The thermal decomposition of UCl42tmu in an oxygen atmosphere was studied. Decomposition of single crystals begins around 180° C and approximates to UCl42tmu(s) + O2(g) → UO2Cl2tmu(s) + tmu(g) + gases and is exothermic (ΔH = −270 ± 5 kJ mol−1). The apparent activation energy for the initial stages (nucleation process) of the reaction was estimated as 362 kJ mol−1. The growth period is described by a one-dimensional diffusion process and the decay period by the contracting-area model.  相似文献   

5.
The absorption spectra of UCl4 single crystals have been measured from 4000 to 30 000 cm−1 at temperatures ranging from 4.2 K to room temperature. The transitions were assigned by comparing the strong absorption features of UCl4 to those of ThCl4:U4+. The energy levels were fitted to a semiempirical Hamiltonian in D2d symmetry.A r.m.s. deviation of 60 cm−1 in the fit to 26 assigned levels was achieved using the following set of parameters (cm−1): F2 = 42 561 ± 235; F4 = 39 440 ± 634; F6 = 24 174 ± 185; ξ = 1805 ± 8; α = 31 ± 1; β = −576 ± 168; (γ = 1200); B20 = −903 ± 151; B40 = 766 ± 220; B44 = −3091 ± 185; B60 = −1619 ± 482; B64 = −308 ± 280.  相似文献   

6.
The gaseous equilibrium S + CF2 = SF + CF was studied over the temperature range 1851 to 2232 K by mass spectrometry, and the derived enthalpy change was used to evaluate the heat of formation of CF ΔH298 = 58.0 ± 2.4 kcal/mol (2.52 ± 0.10 eV), and the dissociation energy D00 (CF) = 130.8 ± 2.4 kcal/mol (5.67 ± 0.10 eV). The new thermochemical data indicate a slightly higher stability for CF than earlier determinations. Direct measurement by electron impact yielded a value of 9.17 ± 0.10 eV for the vertical ionization potential of CF, in agreement with an indirect result obtained from the photodissociative ionization of C2F4.  相似文献   

7.
The intrinsic chemical properties of the gaseous adenine radical cation were examined by using dual cell Fourier transform ion cyclotron resonance mass spectrometry. The adiabatic recombination energy of the radical cation (ionization energy of neutral adenine) was found by bracketing experiments to be 8.55 ± 0.1 eV (at 298 K; earlier literature values range from 8.3 to 8.9 eV). Based on this value, the heat of formation (ΔHf298) of the adenine radical cation is estimated to be 246 ± 3 kcal/mol. The acidity (ΔHacid298) of the adenine radical cation was bracketed to be 221 ± 2 kcal/mol. These thermochemical values suggest that the adenine radical cation reacts with neutral guanine by electron abstraction or proton transfer, with neutral cytosine by proton transfer, and via neither pathway with neutral thymine, molecular water or a sugar moiety of DNA (modeled by tetrahydrofuran). Experimental examination of the gas-phase reactivity of the adenine radical cation revealed a slow deuterium atom abstraction from perdeuterated tetrahydrofuran. Hence, in the absence of a nearby guanine or cytosine, the adenine radical cation may be able to abstract a hydrogen atom from a sugar moiety of DNA.  相似文献   

8.
The bimolecular rate constant for the direct reaction of chlorine atoms with methane was measured at 25°C by using the very-low-pressure-pyrolysis technique. The rate constant was found to be In addition, the ratio k1/k?1 was observed with about 25% accuracy: K1(298) = 1.3 ± 0.3. This gives a heat of formation of the methyl radical ΔH° f 298(CH3) = 35.1 ± 0.15 kcal/mol. A bond dissociation energy BDE (CH3 ? H) = 105.1 ± 0.15 kcal/mol in good agreement with literature values was obtained.  相似文献   

9.
The rate constant of the reaction between Cl atoms and CHF2Br has been measured by chlorine atom resonance fluorescence in a flow reactor at temperatures of 295–368 K and a pressure of ~1.5 Torr. Lining the inner surface of the reactor with F-32L fluoroplastic makes the rate of the heterogeneous loss of chlorine atoms very low (khet ≤ 5 s–1). The rate constant of the reaction is given by the formula k = (4.23 ± 0.13) × 10–12e(–15.56 ± 1.58)/RT cm3 molecule–1 s–1 (with the activation energy in kJ/mol units). The possible role of this reaction in the extinguishing of fires producing high concentrations of chlorine atoms is discussed.  相似文献   

10.
《Chemical physics》1987,117(2):227-235
Time-resolved photoionization mass spectrometry in the millisecond range has been employed to study the reaction C6H5OCH+3 → C6H+6 + CH2O in anisole. Photoionization efficiency (PIE) curves gave a long-time limiting appearance energy value, AE = 10.85 ± 0.05 eV at 298 K. Experimental PIE curves and breakdown graphs at t = 6 μs and 2 ms were compared to those predicted by the statistical theory (RRKM/QET) and by previous photoelectron—photoion coincidence spectrometry results. A sensitivity analysis yielded the following activation parameters: critical energy of activation, E0 = 59.6 ± 0.6 kcal/mol, and entropy of activation, ΔS3(1000 K) = 7.25 ± 2.2 eu.  相似文献   

11.
The heats of the reaction of sodium with ethyl and methyl alcohol were determined by calorimetry. The difference in the standard heats of the formation of triethylarsenite and arsenic trichloride was obtained by calorimetration of the reaction of arsenic trichloride with sodium ethylate, the value of which was −382.42 ± 3.60 kJ/mol. The standard enthalpies of formation were determined from a critical analysis of all data on thermochemistry of trialkylarsenites for the following compounds: triethylarsenite Δf H 298 [(C2H5O)3As(liquid)] = (−704.38 ± 3.85) kJ/mol; trimethylarsenite Δf H 298 [(CH3O)3As(liquid)] = (−599.36 ± 1.88) kJ/mol. The values of standard enthalpies of formation were not adjusted for the following substances in liquid state: arsenic trichloride (−321.96 ± 3.85 kJ/mol), tris-(diethylamido)arsenic(III) As(NEt2)3(liquid) (−129.81 ± 4.41 kJ/mol), tri-n-propylarsenite (−720.61 ± 4.49 kJ/mol), triisopropylarsenite (−756.11 ± 4.65 kJ/mol), tri-n-butylarsenite (−775.11 ± 4.53 kJ/mol), and triisobutylarsenite (−809.71 ± 4.59 kJ/mol). The use of sodium alcoxide solutions for the calorimetration of halogen anhydrides of various acids was demonstrated.  相似文献   

12.
31P, 195Pt and 199Hg NMR spectra of complex (PPh3)2Pt(HgGePh3)(GePh3) (I) have been studied. The spectra at temperatures below ?40°C prove that (I) is a cis-isomer with the square-planar coordination of the Pt atom. The reversibility of temperature dependences of spectra, insensitivity of line shape to the solvent, concentration and presence of free phosphine establish the fluxional behaviour of (I). The activation parameters of the intramolecular rearrangement which is realized, most probably, through a digonal twist, are: Δ298 = 51.5 ± 2.9 kJ/mol, ΔH = 59.3 ± 2.9 kJ/mol, ΔS = 26.2 ± 9.7 J/mol. K.  相似文献   

13.
The kinetics and equilibrium of the gas-phase reaction of CH3CF2Br with I2 were studied spectrophotometrically from 581 to 662°K and determined to be consistent with the following mechanism: A least squares analysis of the kinetic data taken in the initial stages of reaction resulted in log k1 (M?1 · sec?1) = (11.0 ± 0.3) - (27.7 ± 0.8)/θ where θ = 2.303 RT kcal/mol. The error represents one standard deviation. The equilibrium data were subjected to a “third-law” analysis using entropies and heat capacities estimated from group additivity to derive ΔHr° (623°K) = 10.3 ± 0.2 kcal/mol and ΔHrr (298°K) = 10.2 ± 0.2 kcal/mol. The enthalpy change at 298°K was combined with relevant bond dissociation energies to yield DH°(CH3CF2 - Br) = 68.6 ± 1 kcal/mol which is in excellent agreement with the kinetic data assuming that E2 = 0 ± 1 kcal/mol, namely; DH°(CH3CF2 - Br) = 68.6 ± 1.3 kcal/mol. These data also lead to ΔHf°(CH3CF2Br, g, 298°K) = -119.7 ± 1.5 kcal/mol.  相似文献   

14.
The effusion technique with mass spectral recording of ions was employed to investigate the ionic component of molybdenum trifluoride saturated vapour. The equilibrium constants of ion—molecular reactions involving MoF5?, MoF6? and MoOF4? were measured. The following thermodynamic values were obtained from experimental data: MoF4(g) + F?(g) = MoF5?(g),ΔH2980 = ?382.0 ± 20.1 kJ/mole; MoF5(g) + F?(g) = MoF6?(g), ΔH2980 = ?413.4 ± 20.1 kJ/mole; MoOF3(g) + F?(g) = MoOF4?, ΔH2980 = ?418.0 ± 20.5 kJ/mole; EA(MoF5 = 3.6 ± 0.2 eV, EA(MoF6) = 3.6 ± 0.2 eV, EA(MoOF4) = 4.0 ± 0.4 eV. Reported as well as estimated molecular constants were used to calculate thermodynamic functions of some participants of ion—molecular reactions. For MoOF3, BeF3? and Be2F5? vibration frequencies were calculated from the estimated force field.  相似文献   

15.
The kinetics of the gas-phase thermal iodination of hydrogen sulfide by I2 to yield HSI and HI has been investigated in the temperature range 555–595 K. The reaction was found to proceed through an I atom and radical chain mechanism. Analysis of the kinetic data yields log k (l/mol·sec) = (11.1 ± 0.18) – (20.5 ± 0.44)/θ, where θ = 2.303 RT, in kcal/mol. Combining this result with the assumption E?1 = 1 ± 1 kcal/mol and known values for the heat of formation of H2S, I2, and HI, ΔHf,2980(SH) = 33.6 ± 1.1 kcal/mol is obtained. Then one can calculate the dissociation energy of the HS? H bond as 90.5 ± 1.1 kcal/mol with the well-known values for ΔHf,2980 of H and H2S.  相似文献   

16.
The following reaction rate constants of oxygen atoms with iodomethane and chlorine were measured using resonance fluorescence under jet conditions at 298 K: k 1 = (2.4 ± 0.5) × 10–15 and k 2 = (6.9 ± 0.2) × 10−14 cm3/s, respectively.  相似文献   

17.
The dissociation behavior of energy-selected tetraethylsilane, triethylsilane, and diethylsilane photocations is studied using the threshold photoelectron-photoion coincidence (TPEPICO) technique. In the 8–12. 5 eV photon energy range, 0 K dissociation onsets have been measured from the TPEPICO data. The dissociation channels observed include loss of ethane, hydrogen molecule, ethyl radical and hydrogen atom, depending upon the molecular ion under investigation. The thermochemistry of the molecular ions and dissociation fragments is obtained by an analysis that takes into account the kinetics and internal energy distributions of the ions. The various dissociation onsets permit the reevaluation of both neutral and ionic silane thermochemistry. We observed 298-K ethyl group values of 60±10 and 94±10 kJ mol?1 for neutral and ionic silanes, respectively. These values are considerably smaller than the previously reported values of 86 and 130 kJ mol?1, respectively. Finally, a Δ f H ° (298 K) of ?141.5 ± 21 kJ/mol for neutral diethyl silane is derived from the dissociative ionization onset of diethylsilane.  相似文献   

18.
The stepwise complexation of rhenium(V) with N-ethylthiourea has been studied by the potentiometric method in 6 mol/L HCI at 298 K. It has been found that rhenium(V) forms five complex species with this ligand of the following compositions: [ReOLCl4]?, [ReOL2Cl3], [ReOL3Cl2]+, [ReOL4Cl]2+, and [ReOL5]3+. The calculated logarithms of stepwise formation constants of the complexes are the following: logK1 = 4.10 ± 0.05, logK2 = 3.16 ± 0.02, logK3 = 2.61 ± 0.02, logK4 = 2.26 ± 0.02, and logK5 = 1.80 ± 0.02. It has been shown that the introduction of the ethyl radical into the thiourea molecule leads to an increase in the stability of rhenium(V) complexes.  相似文献   

19.
The phase diagram of the pyridine–iron(III) chloride system has been studied for the 223–423 K temperature and 0–56 mass-% concentration ranges using differential thermal analysis (DTA) and solubility techniques. A solid with the highest pyridine content formed in the system was found to be an already known clathrate compound, [FePy3Cl3]·Py. The clathrate melts incongruently at 346.9 ± 0.3 K with the destruction of the host complex: [FePy3Cl3]·Py(solid)=[FePy2Cl3](solid) + liquor. The thermal dissociation of the clathrate with the release of pyridine into the gaseous phase (TGA) occurs in a similar way: [FePy3Cl3]·Py(solid)=[FePy2Cl3](solid) + 2 Py(gas). Thermodynamic parameters of the clathrate dissociation have been determined from the dependence of the pyridine vapour pressure over the clathrate samples versus temperature (tensimetric method). The dependence experiences a change at 327 K indicating a polymorphous transformation occurring at this temperature. For the process ${1 \over 2}[\hbox{FePy}_{3}\hbox{Cl}_{3}]\cdot \hbox{Py}_{\rm (solid)} = {1 \over 2}[\hbox{FePy}_{2}\hbox{Cl}_{3}]_{\rm (solid)} + \hbox{Py}_{\rm (gas)}$ in the range 292–327 K, ΔH $^{0}_{298}$ =70.8 ± 0.8 kJ/mol, ΔS $^{0}_{298}$ =197 ± 3 J/(mol K), ΔG $^{0}_{298}$ =12.2 ± 0.1 kJ/mol; in the range 327–368 K, ΔH $^{0}_{298}$ =44.4 ± 1.3 kJ/mol, ΔS $^{0}_{298}$ =116 ± 4 J/(mol K), ΔG $^{0}_{298}$ =9.9 ± 0.3 kJ/mol.  相似文献   

20.
The equilibrium between fluoral in dichloromethane solution and live condensed liquid polyfluoral has been investigated between 22 and 43°C. Equilibrium monomer concentrations gave: ΔHac°(298 K) = -50-8 ± 2·3 kJ mol?1 and ΔSsc° (298 K) = -142·7 ± 7·4 J K-1 mol-1. With the aid of calibration and monomer vaporization data, thermodynamic values for the polymerization of liquid monomer to liquid polymer were also calculated: ΔHtc° (298 K) = -47 ± 3 kJ mol-1 and ΔS1e° (298 K) = -97 ± 10 J 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