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.
CCSD(T) calculations have been used for identically nucleophilic substitution reactions on N‐haloammonium cation, X? + NH3X+ (X = F, Cl, Br, and I), with comparison of classic anionic SN2 reactions, X? + CH3X. The described SN2 reactions are characterized to a double curve potential, and separated charged reactants proceed to form transition state through a stronger complexation and a charge neutralization process. For title reactions X? + NH3X+, charge distributions, geometries, energy barriers, and their correlations have been investigated. Central barriers ΔE for X? + NH3X+ are found to be lower and lie within a relatively narrow range, decreasing in the following order: Cl (21.1 kJ/mol) > F (19.7 kJ/mol) > Br (10.9 kJ/mol) > I (9.1 kJ/mol). The overall barriers ΔE relative to the reactants are negative for all halogens: ?626.0 kJ/mol (F), ?494.1 kJ/mol (Cl), ?484.9 kJ/mol (Br), and ?458.5 kJ/mol (I). Stability energies of the ion–ion complexes ΔEcomp decrease in the order F (645.6 kJ/mol) > Cl (515.2 kJ/mol) > Br (495.8 kJ/mol) > I (467.6 kJ/mol), and are found to correlate well with halogen Mulliken electronegativities (R2 = 0.972) and proton affinity of halogen anions X? (R2 = 0.996). Based on polarizable continuum model, solvent effects have investigated, which indicates solvents, especially polar and protic solvents lower the complexation energy dramatically, due to dually solvated reactant ions, and even character of double well potential in reactions X? + CH3X has disappeared. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

2.
3.
Laser excited atomic fluorescence-electrothermal atomizer (LEAFS-ETA) was used to study atomization and diffusion mechanisms in a novel diffusive graphite tube atomizer. The atomizer design included a hollow graphite cylinder mounted between two graphite rods which served as electrodes. One of the rods had a small graphite insert with a sampling hollow and could move backwards and forwards. After the sample was introduced into the hollow, the electrodes tightly sealed the graphite cylinder ensuring that the insert was directly in the center of the furnace. The furnace assembly was then heated and the vaporized sample diffused through the hot graphite wall. The atomic fraction of the sample vapor was excited by a laser beam which was directed along the graphite tube surface so that no gap remained between the beam and the tube surface. Fluorescence vs. time profiles for three elements — Cu, Ag and Ni — were obtained within the temperature range of 1400–2600 K. The rate constants of specific processes were measured from the decay portions of the fluorescence signals under the assumption of first-order kinetics. The Arrhenius plots were constructed and the activation energies, Ea were evaluated from their slopes. The plots obtained for Cu and Ag consisted of two linear parts, the corresponding values of Ea were: 195 kJ/mol and 77 kJ/mol for Cu (1550 K < T < 2600 K) and 238 kJ/mol and 97 kJ/mol for Ag (1430 K < T < 2280 K). The Arrhenius plot for Ni was linear within the temperature range of 1770–2530 K resulting in an Ea equal to 161 kJ/mol. The diffusion coefficients were evaluated on the basis of a steady-state diffusion model out of a hollow cylinder. The values for the diffusion coefficients were: 3.7·10−4−2.0·10−3 cm2/s (1750–2600 K) for Cu, 6.5·10−3−1.4·10−3 cm2/s (1750–2280 K) for Ag and 5.6·10−5−1.5·10−3 cm2/s (1770–2530 K) for Ni.  相似文献   

4.
The kinetics of pentoxyl (I) oxidation in aqueous media under the action of hypochlorite ions was studied at pH 8.8 and 273–298 K. The order of the reaction with respect to both participants was found to be one. The temperature dependence of the reaction rate obeyed the Arrhenius law. The reaction activation parameters were found to be E a=11.08 kJ/mol, ΔH =8.73 kJ/mol, ΔS =?200.70 J/(mol K), and ΔG =66.88 kJ/mol. Reaction stoichiometry was studied, the chemical characteristics of the process considered, and a mechanism of the oxidative transformation of I under the action of OCl? suggested.  相似文献   

5.
Nonempirical methods are used to calculate the geometric parameters, the frequencies of normal vibrations, and thermochemical characteristics of ions existing in saturated vapors over sodium bromide and iodide: Na2X+, NaX2, Na3X2+, and Na2X3 (X = Br, I). According to the calculations, Na2X+ and NaX2 triatomic ions have a linear equilibrium configuration of D h symmetry. Pentaatomic ions can exist in the form of three isomers: linear with D h symmetry, planar cyclic with C 2v symmetry, and bipyramidal with D 3h symmetry. At a temperature of ∼1000 K, Na3X2+ and NaX3 pentaatomic ions are shown to be present in vapor mainly in the form of linear isomers. The energies and enthalpies of ion molecular reactions with the participation of the above ions are calculated, and the formation enthalpies of the ions are determined, Δ f H o(0 K): 293±2 kJ/mol (Na2Br+), 354±2 kJ/mol (Na2I+), −536±2 kJ/mol (NaBr2, −458±2 kJ/mol (NaI2, 24±5 kJ/mol (Na3Br2+, 143±5 kJ/mol (Na3I2+, −810±5 kJ/mol (Na2Br3, and −675±5 kJ/mol (Na2I3.  相似文献   

6.
A potential force field has been evaluated for the calculation of the properties of the solid CO-Ar system. The CO·Ar potential energy has been expressed as a sum of the C·Ar and O·Ar interatomic interactions. The (6-exp) Buckingham form of the atom—atom potential, ? = ?Ar?6 + B exp (?αr), has been used (r is the interatomic distance). The values of the A, B and α numerical parameters for the C·Ar and O·Ar potential have been obtained from those for the C·C, O·O, and Ar·Ar potentials using known combining rules. These values are the following: AC·Ar = 3379 kJ/mol A6, BC·Ar = 3.12 × 105 kJ/mol, αC·Ar = 3.493 A?1, AO·Ar = 2737 kJ/mol A6, BO·Ar = 3.28 × 105 kJ/mol, αO·Ar = 3.706 A?1. The three parameters of the Ar·Ar potential function (AAr·Ar = 6554 kJ/mol A6, BAr·Ar = 3.27 × 105 kJ/mol, αAr·Ar = 3.305 A?1) have been fitted to a set of experimental data for the Ar crystal (zero-temperature lattice spacing and energy, and the value of the isothermal compressibility). The CO·Ar potential surface has been calculated showing the most favourable position of an Ar atom near the CO molecule and the orientational dependence of the CO·Ar interactions. The CO·Ar separation distance at the potential minimum and the depth of the potential well are equal to 3.63 A and ?1.321 kJ/mol, respectively. Comparison has been made of the derived Ar·Ar and Co·Ar potential functions with other such functions available in the literature.  相似文献   

7.
The temperature dependence of the Raman spectrum of 1,2-dithiacyclononane (1,2-DTCN) in the SS stretching region has been used to infer the existence of a conformational equilibrium with ΔH0 = 5.0 ± 0.8 kJ/mol. Molecular mechanics calculations predict a (2 2 5)-C2 lowest energy conformation in equilibrium with a (2 3 4) structure. The fully decoupled 13C NMR spectrum at −80°C and the Raman spectra are consistent with this postulate. The temperature dependence of the 1H NMR spectrum of 1,2-DTCN is characteristic of the ring inversion process. A crude lineshape analysis allows us to calculate ΔG0 = 49.0 ± 1.2 kJ/mol.  相似文献   

8.
9.
Data on the static relative permittivity (?s) of 1,2-ethanediol over a temperature range of 243 to 423 K were examined within the framework of the Onsager-Kirkwood-Fröhlich theory. An analysis of the dispersion of the complex relative permittivity ?*(ω) of ethanol using the Davidson-Cole and Vogel-Fulcher-Tammann was performed empirical equations and the Adam-Gibbs molecular-kinetic theory. The correlation factor g exp, relaxation time τ DC, and distribution parameter β were determined. It was established that the parameter β increases from 0.72 to 0.845 as the temperature rises from 243 to 453 K, in contrast to the previous studies, in which it was set temperature-independent. The parameters of the Vogel-Fulcher-Tammann were found to be T 01 = 93.809 K, τ 0VFT1 = 6.96 × 10?14 s, and U ∞VFT1 # = 12.73 kJ/mol. Based on the Adam-Gibbs equation, the number of molecules involved in an elementary event of cooperative structural rearrangement was estimated to be z * = 3.1?2.4 within 161–186 K and z * = 1.63?1.26 within 243–453 K.  相似文献   

10.
The standard enthalpies of formation of liquid and gaseous octachlorotrisilane were estimated, Δf H o (298.15, Si3Cl8, g) = ?1397(9) kJ/mol and Δf H o (298.15, Si3Cl8, l) = ?1447(9) kJ/mol. The decomposition of Si3Cl8 over the temperature range 400–1000 K was studied theoretically.  相似文献   

11.
The enthalpy of the combustion of C60Br24 · 2Br2 has been measured using a rotating-bomb calorimeter as follows: Δ c H 0(C60Br24 · 2Br2, cr) = (?25986 ± 166) kJ/mol. The result has been used to calculate the standard enthalpy of formation, ΔfH 0(C60Br24 · 2Br2, cr) = (2375 ± 166) kJ/mol. The enthalpies of formation of C60Br24 (cr) and dissociation of the C-Br bond have been estimated. The latter value has been compared with enthalpies for the C-X (X = H, F, Cl, Br) bonds in fullerene derivatives and organic compounds.  相似文献   

12.
CdB2X4 spinels—B = Sc, Y, Lu; X = S, Se—were synthesized and their lattice constants were determined. After Mn2+ doping they were investigated by the EPR method over the temperature range 4.2–330 K. The temperature dependence of the hyperfine structure constant, A, was described by the ?imanek-Orbach equation and its parameters were determined. Chemical bonds in the studied spinels were characterized in more detail to obtain a linear dependence of the A constants on the modified covalency parameter.  相似文献   

13.
The data on the dispersion of the permittivity ?*(ω) of 1,2-ethanediol over the temperature range 161–453 K and the frequency range 0.1 Hz–150 GHz were analyzed using the Dissado-Hill cluster model. The relaxation frequency ωp = τ DH ?1 and intra-(n DH) and intercluster (m DH) correlation parameters were calculated. The energy barrier to the libration of molecular axes in clusters was found to be B DH = 2.96 kJ/mol. The apparent enthalpy of activation was determined; it increased from ΔH DH exp # = 22.18 kJ/mol to ΔH DH exp # = 129.19 kJ/mol close to the glass transition temperature. The mean dipole moments $ \bar \mu _c The data on the dispersion of the permittivity ɛ*(ω) of 1,2-ethanediol over the temperature range 161–453 K and the frequency range 0.1 Hz–150 GHz were analyzed using the Dissado-Hill cluster model. The relaxation frequency ωp = τDH−1 and intra-(n DH) and intercluster (m DH) correlation parameters were calculated. The energy barrier to the libration of molecular axes in clusters was found to be B DH = 2.96 kJ/mol. The apparent enthalpy of activation was determined; it increased from ΔH DH exp# = 22.18 kJ/mol to ΔH DH exp# = 129.19 kJ/mol close to the glass transition temperature. The mean dipole moments of 1,2-ethanediol clusters were calculated; they decreased from 162920 to 18.08 D as the temperature increased from 161 to 453 K. According to approximate estimates, the number of 1,2-ethanediol molecules in a cluster /μv decreased from 72405 at 161 K to 8.04 at 453 K (μv is the dipole moment of the molecule in the vacuum), which substantiated the suggestion of the existence of a spatial structure close to the boiling point. Original Russian Text ? N.V. Lifanova, T.M. Usacheva, V.I. Zhuravlev, V.K. Matveev, 2008, published in Zhurnal Fizicheskoi Khimii, 2008, Vol. 82, No. 10, pp. 1973–1981.  相似文献   

14.
The thermodynamic properties and oxygen stoichiometry of Ba2Cu3O5 + δ are studied by means of the electromotive force (EMF) with a fluoride electrolyte, dissolution calorimetry, and thermogravimetry. It is shown that the temperature dependence of the Gibbs energy of the formation of barium cuprate from simple oxides and oxygen in the temperature range of 860–1120 K can be described by the polynomial Δf, ox G °(Ba2Cu3O5 + δ) ± 0.1 (kJ/mol) = ?291.78 + 1.127T ? 0.13207 TlnT (kJ/mol).  相似文献   

15.
Diphenyl(diphenylethynyl)silane ((ph–C≡C)2–Si–ph2) (DPDPES) was synthesized by the Grignard reaction. The corresponding isothermal and non-isothermal cure kinetics of DPDPES were analyzed by using differential scanning calorimetry (DSC), and the molecular structure was characterized by H-NMR. The results showed that all the cure curves were typically sigmoid shape and cure reactions could be described by an autocatalytic kinetic model by isothermal DSC. The kinetic data, for example, activation energy (E) and frequency factor (A), were 119.22 kJ/mol and 4.67 × 107 (s?1), respectively. The non-isothermal DSC analyses showed that E and A were 162.12 kJ/mol and 1.32 × 109 (s?1), respectively, and the reaction order was 0.94. Based on the research work of this paper, it can be said that the cure reaction of DPDPES monomer was of autocatalytic and diffusion-controlled characteristics, and the effect of the diffusion was more evident at low temperature. The cure reaction of DPDPES was a first-order kinetic reaction.  相似文献   

16.
Hydration reactions of protonated and sodiated thiouracils (2-thiouracil, 6-methyl-2-thiouracil, and 4-thiouracil) generated by electrospray ionization have been studied in a gas phase at 10 mbar using a pulsed ion-beam high-pressure mass spectrometer. The thermochemical data, ΔH o n, ΔS o n, and ΔG o n, for the hydrated systems were obtained by equilibrium measurements. The water binding energies of protonated thiouracils, [2SU]H+ and [6Me2SU]H+, were found to be of the order of 51 kJ/mol for the first, and 46 kJ/mol for the second water molecule. For [4SU]H+, these values are 3–4 kJ/mol lower. For sodiated complexes, these energies are similar for all studied systems, and varied between 62 and 68 kJ/mol for the first and between 48 and 51 kJ/mol for the second water molecule. The structural aspects of the precursors for hydrated complexes are discussed in conjunction with available literature data. Graphical Abstract
?  相似文献   

17.
The glass transition temperature (Tg), density, refractive index, Raman scattering spectra, and X-ray photoelectron spectra (XPS) for xZnO-yBi2O3-zB2O3 glasses (x=10-65, y=10-50, z=25-60 mol%) are measured to clarify the bonding and structure features of the glasses with large amounts of ZnO. The average electronic polarizability of oxide ions (αO2−) and optical basicity (Λ) of the glasses estimated using Lorentz-Lorenz equation increase with increasing ZnO or Bi2O3 content, giving the values of αO2−=1.963 Å3 and Λ=0.819 for 60ZnO-10Bi2O3-30B2O3 glass. The formation of BOBi and BOZn bridging bonds in the glass structure is suggested from Raman and XPS spectra. The average single bond strength (BMO) proposed by Dimitrov and Komatsu is applied to the glasses and is calculated using single bond strengths of 150.6 kJ/mol for ZnO bonds in ZnO4 groups, 102.5 kJ/mol for BiO bonds in BiO6 groups, 498 kJ/mol for BO bonds in BO3 groups, and 373 kJ/mol for BO bonds in BO4 groups. Good correlations are observed between Tg and BMO, Λ and BMO, and Tg and Λ, proposing that the average single bond strength is a good parameter for understanding thermal and optical properties of ZnOBi2O3B2O3 glasses.  相似文献   

18.
The vapour pressure of uracil was measured in the temperature range 452–587 K using different techniques and the pressure—temperature equation log P(kPa) = 12.13 ± 0.50 — (6823 ± 210)/T was derived. The thermodynamic functions of gaseous and solid uracil were also evaluated through spectroscopic and calorimetric measurements. The sublimation enthalpy of uracil, ΔH0298 = 131 ± 5 kJ mole?1, was derived from second and third law treatment of the vapour data.  相似文献   

19.
A least-squares linear-Taylor differential-correction technique has been used for the rapid evaluation of thermogravimetric curves obtained during the decomposition of magnesium hydride, iron—titanium hydride and lanthanum—nickel hydride. For magnesium hydride and iron—titanium hydride the Avrami—Erofe'ev equation fits the experimental data, thus indicating that nucleation is the rate-determining step under thermogravimetric conditions. For lanthanum—nickel hydride a combination of the Avrami—Erofe'ev equation and the phase boundary movement equation fits the data up to a fractional decomposition of 0.8. For magnesium hydride decomposition the activation energy E and the pre-exponential factor Z are dependent on the hydrogen pressure (E = 101.2 kJ mole?1 and Z = 8.96 × 107 at 0.30 MPa, while E = 66.3 kJ mole?1 and Z = 4.77 × 107 at 0.11 MPa). For iron—titanium hydride (E = 28.4 kJ mole?1) and lanthanum—nickel hydride (E = 13.4 kJ mole?1) the values are independent of pressure.  相似文献   

20.
A new series of polyimides was synthesized by the condensation of monomers (azomethine‐ether diamine, DA‐1 and DA‐2) with pyromelliticdianhydride (PMDA), 3,4,9,10‐perylenetetracarboxylic dianhydride (PD) and 3,3′4,4′‐benzophenonetetracarboxylic dianhydride (BD). The structural explications of monomers and polyimides was conducted by FT‐IR, 1H NMR and elemental analysis. All polyimides were found soluble in polar aprotic solvents and found to be semicrystalline in nature confirmed by XRD. The inherent viscosities were found in the range of 0.67–0.77 g/dl. %. Average molecular weight (MW) and number average molecular weight (Mn) of the polyimides were found in the range of 5.72 × 105 g/mol–6.58 × 105 g/mol and 3.79 × 105 g/mol 4.11 × 105 g/mol respectively. The polyimides exhibited excellent thermal properties having a glass transition temperature Tg in the range of 230–290°C and the 10% weight loss temperature was above 450°C. The values of thermodynamic parameters, activation energy, enthalpy and entropy fall in the range of 45.2–53.90 kJ/mol, 43.5–52.30 kJ/mol and 0.217 kJ/mol k to 0.261 kJ/mol k respectively. Copyright © 2015 John Wiley & Sons, Ltd.  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