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 46 毫秒
1.
The enthalpies of formation (ΔH f o) for 24 hydrocarbon radicals (R?), mainly polycyclic aromatic radicals with the complex structure, were determined from the published data on bond dissociation energies. The ΔH f o values of the corresponding molecules were calculated, in the majority of cases, by the macroincrement method. Calculations by the group contribution method were performed. Some ΔH f o(R?) values were compared to those calculated by the additive-group method. Calculations were performed, and the conjugation energies of the radicals were discussed. The errors of determination of the ΔH f o(R?) values found were estimated. Due to this work, the database for ΔH f o values of hydrocarbon radicals was increased more than by 25%.  相似文献   

2.
The strain-sensing behaviors of carbon black (CB)/polypropylene (PP) and carbon nanotubes (CNTs)/PP conductive composites prepared by the vacuum-assisted hot compression were studied and compared. When ten extension-retraction cycles were applied, it was found for CB/PP, the value of the maximum responsivity (ΔR/R 0, ΔR—the instantaneous variation of the resistance during the test, R 0—the original resistance) decreased gradually with increasing the cycle number, but it began to rise from the seventh cycle. The value of the min ΔR/R 0 increased during the whole test. While for CNTs/PP, both the values of the max and min ΔR/R 0 decreased rapidly. It is suggested that the different behaviors mainly depend on the distinction in the dimension of the conductive fillers and the preparation technique.  相似文献   

3.
Angle-resolved translational energy spectroscopy has been applied to Cs4I + 3 ions that survived 8 keV collisions with a range of collision gas targets, including inert gases and deuterium. The experimental data comprise values of the translational energy loss ΔTR as a function of the (laboratory-frame) scattering angle θ R for each collision gas under conditions such that single-collision events dominated the scattering. The values of ΔTR increase with θ R, in accordance with very general expectations. However for any value of θ R, the values of ΔTR for helium and deuterium as targets were almost indistinguishable from one another but were at least five to six times larger than those for neon and all other collision gases. These data have been shown to be consistent with theoretical considerations based upon conservation of energy and linear momentum. Theoretical approaches include the simple “elasticlimit” model, which makes no mechanistic assumptions, and a particular “binary-model” theory, which excludes electronic excitation as a possibility. Both theories are consistent with the experimental data and interpret the surprisingly large values of ΔTR for low-mass targets in terms of large recoil energies of the target required to ensure conservation of momentum. The most likely alternative candidate as sink for ΔTR is internal excitation of the target, but this possibility was excluded in the present work by choosing ΔTR values less than the lowest excitation energies of the inert gas targets. Moreover, such an interpretation cannot explain the similar results obtained using helium and deuterium, which were markedly different from those obtained for all other collision gases.  相似文献   

4.
The enthalpy of formation at 298.15 K of the polymer Al13O4(OH)28(H2O)3+8 and an amorphous aluminium trihydroxide gel was studied using an original differential calorimetric method, already developed for adsorption experiments, and aluminium-27 NMR spectroscopy data. ΔHf “Al13” (298.15 K) = ? 602 ± 60.2 kJ mole?1 and ΔHf Al(OH)3 (298.15 K) = ? 51 ± 5 kJ mole?1. Using theoretical values of ΔGR “Al13” and ΔGR Al(OH)3, we calculated ΔGf “Al13” (298.15 K) = ? 13282 kJ mole?1; ΔSf “Al13” (298.15 K) = + 42.2 kJ mole?1; ΔGf Al(OH)3 (298.15 K) = ? 782.5 kJ mole?1; and ΔSf Al(OH)3 (298.15 K) = + 2.4 kJ mole?1.  相似文献   

5.
A new method of group identification was established for selected aliphatic compounds giving homologous series, based upon constant ΔRM and directional coefficient “a” values (the “a” values being tangents of the angle between the course of the RM = f(nc) function and the “x” axis).The results presented concern identification of higher fatty acids and their ethyl esters, amides of higher fatty acids, higher aliphatic amines, ethylalkyl ketones, and dicarboxylic acids.Simultaneously, another method of identifying groups of aliphatic compounds was established, taking advantage of differentiated visualizing effects.  相似文献   

6.
The availability for the first time of detailed rate constants k(V′, R′, T′) (where V′, R′ and T′ are product vibrational, rotational and translational excitation) for the highly exothermic reaction H + F2 → HF(V′, R′) + F has prompted the 3D classical-trajectory study reported here. The potential-energy surface is found to be predominantly repulsive (A ≈ 42%, R ≈ 58%) corresponding to the rather low fractional conversion of reaction energy into vibration ((f′V) = 0.58 from experiment, and 0.56 from theory). In the homologous series of reactions H + X2 (X  F, Cl, Br, I) the percentage of repulsive energy-release decreases for X  Cl, Br, I, but increases from X  F to Cl. It is shown that this cannot be due to charge in mass-combination, but can plausibly be explained by the anomolously short range of interaction between the separating X atoms in the case X  F. It is predicted that the more-forward scattered HF will be more rotationally excited. The form of the cross section function Sr(T) (where T is reagent translation) is analysed. In accordance with the expectation for a strongly exothermic reaction, it is found that Sr(T) rises more steeply than Sr(V) (where V is reagent vibrational energy). The effect on the product energy distribution conforms qualitatively to the “adiabatic” behaviour noted in previous work: ΔT → ΔT′ + ΔR′; ΔV → ΔV′. The explanation is to be found in reaction through more-compressed or more-extended intermediate configurations than are characteristic of room temperature reaction. We note the existence of an amplification effect: (ΔT′ + ΔR′)/ΔT ≈ 2, and ΔV′/ΔV ≈ 2.  相似文献   

7.
Linear correlations have been found between the ΔG0 values of the molecular complexes R1R2R3PO/I2, R1R2SO/I2 and R1R2SeO/I2 and the PO, SO and SeO valence force constants, respectively. The nature of the correlation is determined by the ZO donor bond and not by the donor atom, where Z is P, S or Se. The change on ΔG0 values for an equal change in the ZO valence force constants increases in the order R1R2SO/I2?R1R2R3PO/I2 < R1 R2SeO/I2. Seleninyl complexes with I2 are more stable than the analogous thionyl complexes. From ΔG0fzo correlation deductions can be made about the nature of the ZO donor bond and ΔG0 values can be evaluated from vibrational spectra. A linear correlation exists between the ΔG0 values of corresponding thionyl and seleninyl complexes which is of the same form as the correlation between the valence force constants of analogous thionyl and seleninyl compounds.  相似文献   

8.
The influences of stress ratio (R) on fatigue crack growth (FCG) of thermoset epoxy resin with polyamine hardener were investigated. The FCG growth rates (da/dN) have been correlated by the linear-elastic fracture mechanics parameters (ΔK and Kmax), and nonlinear-elastic fracture mechanics parameter (ΔJ). The effects of R on FCG were observed when the ΔK and ΔJ were used as fracture mechanics parameters for FCG. However, the Kmax successfully characterized FCG under cyclic-dependent condition (FCGs at R = 0.1 and 0.4); but it failed to characterize the FCG under time-dependent condition (FCG at R = 0.7). As a time-dependent fracture mechanics parameter, C* was applied to correlate the time-dependent FCG rate (da/dt). A reasonable agreement was obtained between time-dependent FCG (R = 0.7) and creep crack growth (CCG) results.  相似文献   

9.
The watershed algorithm is the most common method used for peak detection and integration in two-dimensional chromatography. However, the retention time variability in the second dimension may render the algorithm to fail. A study calculating the probabilities of failure of the watershed algorithm was performed. The main objective was to calculate the maximum second-dimension retention time variability, Δ2tR,crit, above which the algorithm fails. Several models to calculate Δ2tR,crit were developed and evaluated: (a) exact model; (b) simplified model and (c) simple-modified model. Model (c) gave the best performance and allowed to deduce an analytical expression for the probability of failure of the watershed algorithm as a function of experimental Δ2tR, modulation time and peak width in the first and second dimensions. It could be demonstrated that the probability of failure of the watershed algorithm under normal conditions in GC × GC is around 15–20%. Small changes of Δ2tR, modulation time and/or peak width in the first and second dimension could induce subtle changes in the probability of failure of the watershed algorithm. Theoretical equations were verified with experimental results from a diesel sample injected in GC × GC and were found to be in good agreement with the experiments.  相似文献   

10.
Optically active 1-fluoroindan-1-carboxylic acid (FICA) was designed and prepared as its methyl ester for determining the absolute configuration of chiral molecules by both 1H and 19F NMR spectroscopies. Enantiomerically pure isomers of FICA methyl esters (FICA Me esters) were obtained by chromatographic separation using HPLC with a Daicel Chiralcel OJ-H column. The absolute configuration of the (+)-FICA Me ester was deduced to be (S) by X-ray crystallographic analysis of the (+)-FICA amide of (R)-α-phenethylamine. Both enantiomers were derived to the diastereomeric esters of chiral secondary alcohols by an ester exchange reaction. In the 1H NMR spectra, the signs of ΔδH (δR ? δS) were consistent on each side of the FICA molecular plane. Therefore, the concept of the modified Mosher’s method could be successfully applied to the FICA-based procedure. Moreover, the consistency in the signs of ΔδF (δR ? δS) values suggests that the FICA method would be reliable in assigning the absolute configurations of secondary alcohols based on 19F NMR spectroscopy.  相似文献   

11.
Thermal and volume parameters of the twin charge-ordering Verwey transitions in RBaFe2O5+w (R=a rare-earth element) are summarized as a function of R and w. Their determination is exemplified for case R = Dy, for which also synthesis conditions, phase relations, and refined crystal-structure data for the valence-mixed (Fe2.5+) and charge-ordered (Fe2+ and Fe3+) phases are reported. Data for the R=Nd, Sm, Eu and Gd variants with wide ranges of oxygen non-stoichiometry suggest that increasing w decreases ΔS and the temperature of the transition in a manner that is similar to a behavior under increasing concentration of an ideal solution of RBaFe2O6 in RBaFe2O5. Thermal parameters of the transition for the ideal mixed-valence composition RBaFe2O5.000 are estimated from such compositional dependences, varying reasonably smoothly as a function of R (radius, electronegativity, polarizability). Parameter ΔV is the only one that follows the structural discontinuity between the charge-ordered R = Nd and variants with smaller trivalent R ions. The ordering of the dxz orbitals of the Fe2+ ions is thus being achieved at a cost of lowering the symmetry when the R size becomes unfavorably large. A definition of the Verwey transition as a first-order orbital ordering of a valence-mixed phase is suggested.  相似文献   

12.
The heat of solution (ΔH s ) of several homologous series of linear and monofunctional organic compounds (n-alkanes,n-alcohols,n-aldehyds and esters) in polar stationary phase (Carbowax 1540) can be expressed by ΔH s =a+b·n R +c·C G , wherea is a constant,n R is the number of carbon atoms not belonging to the functional group in the molecule andC G represents the contribution of each functional group to the ΔH s value. The accuracy of the ΔH s values calculated by this equation is sufficient for all the compounds tested (average relative error=1.92%).C G can be calculated from the plots ΔH s versusn R .  相似文献   

13.
《Tetrahedron: Asymmetry》1998,9(4):563-574
Homochiral crown ether (S,S)-1 containing 1-naphthyl groups as chiral barriers together with the phenol moiety was prepared by using (S)-3 as a chiral subunit which was resolved in enantiomerically pure form by lipase-catalyzed enantioselective acylation of (±)-3. Homochiral phenolic crown ether (S,S)-2, containing phenyl groups as chiral barriers, was also prepared from (S)-5 which was derived from (S)-mandelic acid. The association constants for their complexes with chiral amines in CHCl3 were determined at various temperatures by the UV–visible spectroscopic method demonstrating that the crown ethers (S,S)-1 and (S,S)-2 displayed the large ΔRSΔG values of 6.2 and 6.4 kJ mol−1, respectively, towards the amine 21 at 15°C. Thermodynamic parameters for complex formation were also determined and a linear correlation between TΔRSΔS and ΔRSΔH values was observed.  相似文献   

14.
The correlation factors Rp and Rs are determined for the dielectric polarization and the non-linear dielectric effect in dilute solutions of nitrobenzene in benzene. Assuming dipole association of the nitrobenzene molecules to be restricted to dimerization, we determined the concentration x2 of dimers, their dipole moment μdim, and the equilibrium constant Kx and Gibbs energy ΔG of the dimerization process.  相似文献   

15.
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   

16.
For measuring solution enthalpies of the strong hygroscopic double chlorides, an isoperibol solution calorimeter was built. Samples of 2–5 g could be solved to a molar dilution 1 : 3000. The temperature difference between reaction vessel and thermostat was measured by a thermopile; the temperature of the thermostat was constant to 2 · 10?4°C. From the molar enthalpies of solution (ΔHL), enthalpies for the reactions nACl + MCl2 = AsMCl(s + 2) were calculated: ΔHPR = ? ΔHL (double chloride) + ΔHL(n Cl) + ΔHL(MCL2) These values are relatively small: about ? 50 kJ for the Cs-compounds, nearly zero for the K- and Na-compounds.  相似文献   

17.
Theoretical study on hydration of carbonyl compounds has been done at B3LYP/6-31++G** and MP2/6-31++G** levels. The variations in ΔG hyd and hydration constants are explored in terms of medium effect, substituent effect, and hydrogen bonding abilities of carbonyl compounds and their hydrated products. The dielectric of medium decreases the ΔG hyd values thereby favoring the process. The presence of electron-releasing substituents at the carbonyl carbon disfavors the hydration process, while that of electron-withdrawing substituents favor the process. Hydrogen bonding interactions stabilize the product to a larger extent than the carbonyl molecules, thereby favoring the hydration process. Linear correlation between the calculated log K hyd values and the experimental values is seen in case of specific interactions with water (R = 0.976) than in the case without those interactions (R = 0.955).  相似文献   

18.
Four new compounds composed of chiral complex cations and anions: Δ-[Fe(bpy)3] Λ-[Fe(bpy)3][V4O8((2R,3R)-tart)2]·12H2O (1), Δ-[Fe(bpy)3]2Λ-[Fe(bpy)3]2[V4O8((2R,3R)-tart)2][V4O8((2S,3S)-tart)2]·24H2O (2), Δ-[Ni(bpy)3]Λ-[Ni(bpy)3][V4O8((2R,3R)-tart)2]·12H2O (3) and Δ-[Ni(bpy)3]2Λ-[Ni(bpy)3]2[V4O8((2R,3R)-tart)2][V4O8((2S,3S)-tart)2]·24H2O (4) have been prepared. The compounds have been characterized by spectral methods, and their thermal decomposition was studied by simultaneous DTA, TG measurements. The final products after dynamic decomposition and additional heating were Fe2V4O13 for 1 and 2 and Ni(VO3)2 for 3 and 4. The crystal structures determined for 1, 2 and 4 have evidenced that 1 is “hemiracemic” and 2 and 4 are “fully racemic” compounds.  相似文献   

19.
20.
《Polyhedron》1999,18(6):905-908
The aluminum alkoxides [MeClAlOEt]3 (1), [Et2AlOMe]3 (2), [Me2AlOEt]3 (3), and [EtClAlOEt]3 (4) were investigated by 1H NMR spectroscopy to study the o-dichlorobenzene solution equilibrium: 2 [R2AlOR′]3⇌3 [R2AlOR′]2. The complexes are shown to exist primarily as trimers at room temperature, but increasing concentrations of the dimeric form are observed at higher temperatures. Equilibrium constants, ΔH, and ΔS were determined for the trimer–dimer equilibrium. Values of ΔH for the conversion of 2 moles of trimer are 63(4), 78(1), 85.0(8), and 99(6) kJ for 1, 2, 3, and 4, respectively. The corresponding values of ΔS are 142(7), 184(3), 218(2), and 265(17) J/K, respectively. Thermodynamic parameters are compared with those reported for [Me2AlOPrn]3 and [Me2AlOPh]3. The characterization of [EtClAlOMe]3 is also reported.  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