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 15 毫秒
1.
《Chemical physics letters》1986,127(4):343-346
In this work we use a complete surface hopping quasiclassical trajectory method to determine cross sections for the reactions H2+ + H2 → H3+ + H and the isotopic variants (H2+ + D2 and D2+ + H2). Initial translational energies ranged between 0.5 and 6 eV. The vibrational quantum number (v+) of the charged diatom is either 0 or 3. Comparing these results with our previous results with a partial treatment of surface hopping, we find essentially no change for v+ = 0 and reductions in cross sections of up to 30% for v+ = 3 trajectories.  相似文献   

2.
《Fluid Phase Equilibria》1998,152(2):277-282
Excess molar volumes VmE have been measured using a dilatometric technique for mixtures of cyclohexanone (C6H10O) with trichloromethane (CHCl3), 1,2-dichloroethane (CH2ClCH2Cl), trichloroethene (CHClCCl2), 1,1,1-trichloroethane (CCl3CH3), and cyclohexane (c-C6H12) at T=308.15 K, and for cyclohexanone+dichloromethane (CH2Cl2) at T=303.15 K. Throughout the entire range of the mole fraction χ of C6H10O, VmE has been found to be positive for χ C6H10O+(1−χ)c-C6H12, and negative for χ C6H10O+(1−χ)CH2Cl2, χ C6H10O+(1−χ)CHClCCl2, χ C6H10O+(1−χ)CHCl3, and χ C6H10O+(1−χ) CCl3CH3. For χ C6H10O+(1−χ)CH2ClCH2Cl, VmE has been found to be positive at lower values of χ and negative at high values of χ, with inversion of sign from positive to negative values of VmE for this system occurring at χ∼0.78. Values of VmE for the various systems have been fitted by the method of least squares with smoothing equation, and have been discussed from the viewpoint of the existence specific interactions between the components.  相似文献   

3.
Specific ion/molecule reactions are demonstrated that distinguish the structures of the following isomeric organosilylenium ions: Si(CH3) 3 + and SiH(CH3)(C2H5)+; Si(CH3)2(C2H5)+ and SiH(C2H5) 2 + ; and Si(CH3)2(i?C3H7)+, Si(CH3)2(n?C3H7)+, Si(CH3)(C2H5) 2 + , and Si(CH3)3(π?C2H4)+. Both methanol and isotopically labeled ethene yield structure-specific reactions with these ions. Methanol reacts with alkylsilylenium ions by competitive elimination of a corresponding alkane or dehydrogenation and yields a methoxysilylenium ion. Isotopically labeled ethene reacts specifically with alkylsilylenium ions containing a two-carbon or larger alkyl substituent by displacement of the corresponding olefin and yields an ethylsilylenium ion. Methanol reactions were found to be efficient for all systems, whereas isotopically labeled ethene reaction efficiencies were quite variable, with dialkylsilylenium ions reacting rapidly and trialkylsilylenium ions reacting much more slowly. Mechanisms for these reactions and differences in the kinetics are discussed.  相似文献   

4.
Quasi-classical dynamic threshold energies have been determined for two important reactive transitions in the F + H2 reaction by performing extensive three-dimensional trajectory calculations for the corresponding reverse reactions. It is found that the energetic and dynamic thresholds for the reaction F + H2(υ = 0,j) → HF(υ = 2,j = 6) + H are the same, whereas the latter threshold is approximately 0.08 eV greater than the former one for the reaction F + H2(υ = 0,j) → HF(υ = 3,j=1) + H. These results are in good agreement with the corresponding semi-classical threshold results which are also reported. The relationship of these quasi-classical-reverse results to experimentally measured quantities is discussed.  相似文献   

5.
Chand, A., McQuillan, A.R. and Fenby, D.V., 1979. Thermodynamic study of systems with lower critical solution temperatures: H2O + (C2H5)3N, D2O + (C2H5)3N. Fluid Phase Equilibria, 2: 263–274.Molar excess enthalpies and molar excess volumes are reported for the systems H2O + (C2H5)3N and D2O + (C2H5)3N at temperatures below and above their lower critical solution temperatures. The molar excess enthalpies are slightly less exothermic for the D2O system. The molar excess volumes of the H2O and D2O systems are within experimental error of one another. Compositions of conjugate solutions estimated from the calorimetric and volumetric measurements agree with those obtained from published liquid—liquid phase diagrams.  相似文献   

6.
《Fluid Phase Equilibria》1996,126(2):233-239
Excess molar volumes at 298.15 K and atmospheric pressure were measured for {x1 CH3CO2(CH2)3CH3 + x2 C10H22 + (1 − x1x2) Cl(CH2)3CH3} and the corresponding binary mixtures, with an Anton Paar densimeter. All the experimental values were compared with the results obtained by different prediction methods.  相似文献   

7.
Fourier transform ion cyclotron resonance (FTICR) mass spectrometry has been used to examine the reactions of Sc(OCD3)2+ with water, ethanol, and 1-propanol. Sigma-bond metathesis resulting in the elimination of CD3OH is the initial reaction observed, with further solvation of the metal center and subsequent elimination of hydrogen occurring as additional reaction channels. These processes are facile at room temperature and involve little or no activation energy. Measured equilibrium constants for the reaction Sc(OCD3)2+ +ROH ⇌ CD3OScOR+ +CD3OH with R =H, ethyl, and n-propyl are 0.013 ±0.004, 0.5 ±0.15, and 0.7 ±0.2, respectively. For the reaction ROScOCD3+ +ROH ⇌ Sc(OR)2+ +CD3OH with R =H and ethyl the measured equilibrium constants are 0.013 ±0.004 and 0.3 ±0.1, respectively. ΔS is estimated for these processes using theoretical calculations and statistical thermodynamics, and in conjunction with the measured equilibrium constants we have evaluated ΔH for these reactions and the relative and absolute bond strengths of the Sc+–OR bonds, R =H, methyl, ethyl, and n-propyl. The relative bond strengths, D298o(CD3OSc+–OR)–D298o(CD3OSc+–OCD3), for R =H, methyl, ethyl, and n-propyl are +11.9, 0, −0.1, and −1.4 kcal mol−1, respectively. The absolute bond strengths for HOSc+–OCD3, CD3OSc+–OCD3, CD3OSc+–OC2H5, CD3OSc+–OCH2CH2CH3, and H5C2OSc+–OC2H5 are 115.0, 115.0, 114.9, 113.6, and 114.7 kcal mol−1, respectively. Theoretical calculations with an LAV3P1 ECP basis set at the level of localized second-order Møller–Plesset perturbation theory were performed to evaluate ΔS and ΔG for the specific equilibria Sc(OH)2+ +CD3OH ⇌ CD3OScOH +H2O, CD3OScOH +CD3OH ⇌ Sc(OCD3)2+ +H2O, and Sc(OCD3)2+ +C2H5OH ⇌ CD3OScOC2H5+ +CD3OH. The theoretically determined ΔG values agree reasonably well with the experimentally determined ΔG values. In accordance with earlier theoretical predictions, these metathesis reactions are consistent with an allowed four-center mechanism similar to that of a 2σ +2σ cycloaddition.  相似文献   

8.
Boiling temperature measurements have been made at ambient pressure for saturated ternary solutions of NaCl + KNO3 + H2O, NaNO3 + KNO3 + H2O, and NaCl + Ca(NO3)2 + H2O over the full composition range, along with those of the single salt systems. Boiling temperatures were also measured for the four component NaCl + NaNO3 + KNO3 + H2O and five component NaCl + NaNO3 + KNO3 + Ca(NO3)2 + H2O mixtures, where the solute mole fraction of Ca(NO3)2, x{Ca(NO3)2}, was varied between 0 and 0.25. The maximum boiling temperature found for the NaCl + KNO3 + H2O system is ≈134.9 C; for the NaNO3 + KNO3 + H2O system is ≈165.1 C at x(NaNO3) ≈ 0.46 and x(KNO3) ≈ 0.54; and for the NaCl + Ca(NO3)2 + H2O system is 164.7 ± 0.6 C at x{NaCl} ≈ 0.25 and x{Ca(NO3)2} ≈ 0.75. The NaCl + NaNO3 + KNO3 + Ca(NO3)2 + H2O system forms molten salts below their maximum boiling temperatures and the temperatures corresponding to the cessation of boiling (dry-out temperatures) of these liquid mixtures were determined. These dry-out temperatures range from ≈300 C when x{Ca(NO3)2} = 0 to ≥ 400 C when x{Ca(NO3)2} = 0.20 and 0.25. Mutual deliquescence/efflorescence relative humidity (MDRH/MERH) measurements were also made for the NaNO3 + KNO3 and NaCl + NaNO3 + KNO3 salt mixture from 120 to 180 C at ambient pressure. The NaNO3 + KNO3 salt mixture has a MDRH of 26.4% at 120 C and 20.0% at 150 C. This salt mixture also absorbs water at 180 C, which is higher than expected from the boiling temperature experiments. The NaCl + NaNO3 + KNO3 salt mixture was found to have a MDRH of 25.9% at 120 C and 10.5% at 180 C. The investigated mixture compositions correspond to some of the major mineral assemblages that are predicted to control brine composition due to the deliquescence of salts formed in dust deposited on waste canisters in the proposed nuclear repository at Yucca Mountain, Nevada.  相似文献   

9.
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   

10.
The chelate compounds K[Fe(hyc)3] and N2H5[Fe(hyc)3]·H2O (hyc = N2H3COO) were studied by the Mössbauer effect of 57Fe at various temperatures. At room temperature the quadrupole splitting parameter is 2.77 mm/sec for K[Fe(hyc)3] and 2.35 mm/sec for N2H5[Fe(hyc)3]·H2O, and the center shift is 1.08 mm/sec for both compounds. The temperature dependences of the quadrupole parameters yielded the crystal field splittings of the 5T2g levels of the Fe2+ ions which indicate large trigonal distortion of the Fe(hyc)3 anion. Using a molecular crystal-like treatment of the ferrous ion vibrations the temperature dependence of the recoilless fraction gave an effective Debye temperature ΘD = 71°K for K[Fe(hyc)3] and ΘD = 90°K for N2H5[Fe(hyc)3]·H2O. No evidence for magnetic ordering was found down to 4.5°K in either compound.  相似文献   

11.
This article studies the solubility, Hansen solubility parameters (HSPs), and thermodynamic behavior of a naturally-derived bioactive thymoquinone (TQ) in different binary combinations of isopropanol (IPA) and water (H2O). The mole fraction solubilities (x3) of TQ in various (IPA + H2O) compositions are measured at 298.2–318.2 K and 0.1 MPa. The HSPs of TQ, neat IPA, neat H2O, and binary (IPA + H2O) compositions free of TQ are also determined. The x3 data of TQ are regressed by van’t Hoff, Apelblat, Yalkowsky–Roseman, Buchowski–Ksiazczak λh, Jouyban–Acree, and Jouyban–Acree–van’t Hoff models. The maximum and minimum x3 values of TQ are recorded in neat IPA (7.63 × 10−2 at 318.2 K) and neat H2O (8.25 × 10−5 at 298.2 K), respectively. The solubility of TQ is recorded as increasing with the rise in temperature and IPA mass fraction in all (IPA + H2O) mixtures, including pure IPA and pure H2O. The HSP of TQ is similar to that of pure IPA, suggesting the great potential of IPA in TQ solubilization. The maximum molecular solute-solvent interactions are found in TQ-IPA compared to TQ-H2O. A thermodynamic study indicates an endothermic and entropy-driven dissolution of TQ in all (IPA + H2O) mixtures, including pure IPA and pure H2O.  相似文献   

12.
Infrared chemiluminescence from HF and HCl has been observed and yielded vibrational and rotational population distributions for the reactions F + HBr, F + H2Se, and Cl + H2Se. Evaluation of the spectra recorded by a commercial Fourier-transform spectrometer under low-flow conditions gave the following relative vibrational populations: for F ? HBr. Nυ = 1 : Nυ = 2 : Nυ = 3 : Nυ = 4 = 0.45 : 0.31 : 0.13 : 0.11: for F + H2Se, Nυ = 1 : Nυ = 2 : Nυ = 3 : Nυ = 4 : Nυ = 5 = 0.29 : 0.35 : 0.24 : 0.09 : 0.03: for Cl + H2Se, Nυ = 1 : Nυ = 2 : Nυ = 3 = 0.40 : 0.51 : 0.09. All three vibrational surprisal plots show a significant deviation from linearity. Neglecting the contributions from Nυ = 0, the total energy is partitioned into vibration and rotation as follows: 〈fV〉 = 0.49 and 〈fR〉 = 0.09 for F + HBr, 〈fV〉 = 0.41 and 〈fR〉 = 0.07 for F + H2Se, 〈fV〉 = 0.53 and 〈fR〉 = 0.10 for Cl + H2Se. Inclusion of estimates for Nυ = 0 gives the more realistic values 〈fV〉 = 0.24, 0.34, and 0.49 respectively. Whereas 9 ± 3% of the collisions between F + HBr yield Br in the excited 2P12 state, no rovibrationally excited HSe fragments were detected in the two other systems. Consistent values for the bond dissociation energy D00(HSeH) = 329 ± 5 kJ/mol and the enthalpy of formation ΔH100 (HSe) = 137 ± 5 kJ/mol are derived from the highest observed HCl and HF levels.  相似文献   

13.
14.
A flow microcalorimeter of the Picker design was used to measure excess molar enthalpies HE at 298.15 K as a function of mole fraction χ1 for several mixtures belonging to series I: {χ11,2,4-C6H3(CH3)3 + χ2n-CH2ℓ+2}, and series II: {χ11-C10H7CH3 + χ2n-CH2ℓ+2}. The chain length ℓ of the n-alkanes ranged between 7 and 16. 1,2,4-trimethylbenzene and 1-methylnaphthalene have about the same size and shape as the previously investigated chloro derivatives 1,2,4-C6H3Cl3 and 1-C10H7Cl but a much smaller reduced dipole moment. The calorimeter was used in the discontinuous mode. A plot of HEmax (i.e., the maximum value of HE with respect to composition) against ℓ for series I shows a shallow minimum around ℓ = 11 with HEmax (ℓ = 11) ≈ 250 J mol−1, whereas HEmax for series II decreases over the whole range 7 ⩽ ℓ ⩽ 16: HEmax (ℓ = 7) ≈ 760 J mol−1, and HEmax (ℓ = 16) ≈ 595 J mol−1. The corresponding enthalpic interaction parameters h12, calculated from zeroth-order KGB (Kehiaian-Guggenheim-Barker) theory, decrease with increasing ℓ, and the rate of decrease, dh12/dℓ, diminishes for larger chain lengths.For three mixtures belonging to series I (ℓ = 7, 10, 14), excess molar volumes VE and excess molar heat capacities CEP at constant pressure were mesured at the same temperature. VE was determined with a vibrating-tube densimeter (flow conditions), and CEP was obtained with another type of flow calorimeter (stepwise procedure). VE1 = 0.5)/(cm3 mol−1) = −0.207 for ℓ = 7, 0.060 for ℓ = 10, and 0.145 for ℓ = 14. The corresponding values for CEP x1 = 0.5)/(J K−1 mol−2) are 0.32, 0.66 and −0.09. Thus the chain length dependence of the excess molar heat capacity is qualitatively similar to that observed for the series with the homomorphic chloro derivative, (1,2,4-C6H3Cl3 + n-CH2ℓ+2), and to that of (1-C10H7Cl+n-CH2ℓ+2).  相似文献   

15.
The aqua rhenium oxocomplex [ReO(OH)(H2O)4]2? (1) has been prepared and characterized by spectroscopy, thermogravimetry, and elemental analysis and its reactivity towards triphenylphosphine has been evaluated. Complex (1) acts as a catalyst precursor in the presence of molecular oxygen for the oxidation of PPh3 to OPPh3. This proceeds through complex intermediates like [Re(PPh3)n]3+ (2), and [ReO(PPh3)n]3+ (3). The newly prepared complex (1) was also employed as catalyst for catalytic oxidation of cyclohexane. The geometry of [ReO(OH)(H2O)4]2? has also been optimized in the singlet state by the DFT method with B3LYP level of theory.  相似文献   

16.
Excess molar enthalpies HmE of triethylamine + ethylbenzene, + n-propylbenzene, + n-butylbenzene, + isopropylbenzene, and + isobutylbenzene were measured over the entire composition range at 303.15 K with an LKB flow microcalorimeter. HmE values are positive and decrease with increasing chain length of the alkylbenzene.  相似文献   

17.
We have carried out absorption and photoluminescence spectra of six newly prepared TmF3 (2 M%) doped fluorophosphate glasses of composition: (NaPO3)6BaF2ZnF2RF where R = Li, Na, K, (LiNa), (NaK) and (KLi). By evaluating the Judd—Ofelt intensity parameters for the measured absorption levels, the recorded photoluminescence spectra have been analysed to estimate the fluorescent radiative properties (A, AT, βR) and the stimulated emission cross-sections (σEP) of the following luminescence transitions:1G43H6 (λ=451 nm)1D23H4 (λ=468 nm)3P03F2 (λ=482 nm)1D23H5 (λ=492 nm)These have been determined in order to understand the alkali effects on the Tm3+-optical glasses studied.  相似文献   

18.
19.
Photoelectron energy and angular distributions are measured for the 2+1 multiphoton ionization process H2 X1Σg+ (ν = 0,J) + 2hv → E,F1Σg+E,JE = J) + hν → H2+X2Σg++) + e?, for νE = 0, 1, or 2 and for JE = 0 or 1 of the inner well of the double-minimum E,F state. Although a strong preference is found for ν+ = νE, the detailed H2+ vibrational distribution does not exhibit Franck-Condon behavior, and the photoelectron angular distributions vary markedly with both the JE value of the intermediate state and the ν+ value of the ion.  相似文献   

20.
The interaction in the system Bi(NO3)3-K3HCit-(H2O + glycerol) was studied by chemical analysis and pH titration over a wide range of ratios between the initial components. Solid phases of the compositions Bi(OH)3?3x HCit x · nH2O and KBiCit · H2O were isolated and investigated by X-ray powder diffraction and differential thermal analyses and IR spectroscopy. The effect of the chemical nature of the alkali metal on the solubility in the system under consideration was demonstrated.  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