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 62 毫秒
1.
The addition of electrons to form gas-phase multiply charged anions (MCAs) normally requires sophisticated experiments or calculations.In this work, the factors stabilizing the MCAs, the maximum electron uptake of gas-phase molecules, X, and the electronic stability of MCAs XQ-, are discussed. The drawbacks encountered when applying computational and/or conceptual density functional theory (DFT) to MCAs are highlighted. We develop and test a different model based on the valence-state concept. As in DFT, the electronic energy, E(N, vex), is a continuous function of the average electron number, N, and the external potential, vex, of the nuclei. The valence-state-parabola is a second-order polynomial that allows extending E(N, vex) to dianions and higher MCAs. The model expresses the maximum electron acceptance, Qmax, and the higher electron affinities, AQ, as simple functions of the first electron affinity, A1, and the ionization energy, I, of the "ancestor" system. Thus, the maximum electron acceptance is Qmax, calc = 1 + 12A1/7(I -A1). The ground-state parabola model of the conceptual DFT yields approximately half of this value, and it is termed Qmax, GS = ${}^{1}\!\!\diagup\!\!{}_{2}\; $ + A1/(I -A1). A large variety of molecules are evaluated including fullerenes, metal clusters, super-pnictogens, super-halogens (OF3), super-alkali species (OLi3), and neutral or charged transition-metal complexes, ABmLn0/+/-. The calculated second electron affinity A2, calc = A1-(7/12)(I -A1) is linearly correlated to the literature references A2, lit with a correlation coefficient R = 0.998. A2 or A3 values are predicted for further 24 species. The appearance sizes, nap3-, of triply charged anionic clusters and fullerenes are calculated in agreement with the literature.  相似文献   

2.
The e.m.f. of the galvanic cells Pt,C,Te(l),NiTeO3,NiO/15 YSZ/O2 (Po2 = 0.21 atm),Pt and Pt,C,NiTeO3,Ni3TeO6,NiO/15 YSZ/O2 (Po2 = 0.21 atm),Pt (where 15 YSZ=15 mass% yttria-stabilized zirconia) was measured over the ranges 833–1104 K and 624–964 K respectively, and could be represented by the least-squares expressions E(1)±1.48 (mV) = 888.72 − 0.504277 (K) and E(II) ±4.21 (mV) = 895.26 − 0.81543T (K).

After correcting for the standard state of oxygen in the air reference electrode, and by combining with the standard Gibbs energies of formation of NiO and TeO2 from the literature, the following expressions could be derived for the ΔG°f of NiTeO3 and Ni3TeO6: ΔGf°(NiTeO3) ± 2.03 (kJ mol−1) = −577.30 + 0.26692T (K) and ΔG°f(Ni3TeO6)±2.54 (kJ mol−1) = −1218.66 + 0.58837T (K).  相似文献   


3.
The monolayer behavior of three mixed systems of dipalmitoyl phosphatidyl choline (DPPC) with sterols; cholesterol (Ch), stigmasterol (Stig), and cholestanol (Chsta) formed at the interface of air/water (phosphate buffer solution at 7.4 with addition of NaCl) was investigated in terms of surface pressure (π) and molecular occupation surface area (A) relation. A series of πA curves at every 0.1 mol fraction of each sterol for the three combinations of mixed systems were obtained at 25.0 °C.

On the basis of the πA curves, the additivity rule in regard to A versus sterol mole fraction (Xst) was examined at discrete surface pressures such as 5, 10, 15, 20, 25, 30 mN m−1, and then from the obtained AXst curves the partial molecular areas (PMA) were determined. The AXst relation exhibited a marked negative deviation from ideal mixing in the pressure range below 10 mN m−1, i.e. in the expanded liquid film region (below the transition pressure of DPPC).

The PMA of Ch at π=5 mN m−1, for example, was found to be conspicuously negative in the range of XCh=0–0.2 (about −0.4 nm2 per molecule) and slightly positive (ca. 0.1 nm2 per molecule) in the range XCh=0.2 to 0.4. Above XCh=0.5, Ch’s PMA was almost the same as the surface area of pure Ch, while DPPC’s PMA was reduced to 60% of that of the pure system.

Excess Gibbs energy (ΔG(ex)) as a function of Xst was estimated at different pressures. Applying the regular solution theory to thermodynamic analysis of ΔG(ex), the activity coefficients (f1 and f2) of DPPC and the respective sterols as well as the interaction parameter (Ip) in the mixed film phase were evaluated; the results showed a marked dependence on Xst.

Compressibility Cs and elasticity Cs−1 were also examined. These physical parameters directly reflected the mechanical strength of formed monolayer film.

Phase diagrams plotting the collapse pressure (πc) against Xst were constructed, and the πc versus Xst curves were examined for the respective mixed systems in comparison with the simulated curves of ideal mixing based on the Joos equation.

Comparing the monolayer behavior of the three mixed systems, little remarkable difference was found in regard to various aspects. In common among the three combinations, the mole fraction dependence in monolayer properties was classified into three ranges: 0<Xst<0.2, 0.2<Xst<0.4 and 0.5<Xst<1. How the difference in the chemical structure of the sterols influenced the properties was examined in detail.  相似文献   


4.
Thermogravimetric analysis of hydroxyl terminated polybutadiene (HTPB) and its fractions of different molecular weights separated by preparative GPC shows two major stages of weight loss of different nature in a nitrogen atmosphere. The first stage is primarily depolymerisation, cyclisation and crosslinking of molecules and the second stage is mainly the decomposition of the residue from the first stage. The kinetic parameters, viz. activation energy E and pre-exponential factor A using four different non-isothermal integral equations show a systematic increase with increase in molecular weight for the first stage, whereas for the second stage, the effect of molecular weight on E and A values is not prominent. The increase in E and A values for the first stage is attributed to the formation of greater number of cyclised and crosslinked products from molecules of higher dimensions. Quantitative correlations between the kinetic constants and the molecular weight parameters were derived for the first stage as a quadratic curve following the equation: E or ln A = K1K2/M (where K1 and K2 are empirical constants whose values are different for the different molecular weight averages, viz. Mn, Mw and Mz and for the different equations).  相似文献   

5.
The dependence of the Stern potential, ψ1, of glass samples on the distance between these, H, has been theoretically calculated, while taking into account the Stern isotherm and the electroneutrality equation. Comparison of the theoretical dependences ψ1(C)H→∞ with those previously experimentally obtained enables one to calculate the energy of adsorption of OH ions on glass and, further, the dependence ψ1(H). It has been shown that for pH 4–6 and CKCl = 10-2-10-5 mol/L, the value of ψ1 practically does not depend on H. The result obtained was used to calculate theoretically the ionic-electrostatic forces and to compute (from the experimental values of the interaction forces) structural forces Us(H). The dependence thus obtained, Us(H), is of exponential character.  相似文献   

6.
The calculations reported here assign a charge qN = −0.52 electron units to the terminal nitrogen atoms in the azide ion and a value of 141.9 kJ mole−1 to the enthalpy of formation of the gaseous azide ion, ΔHf0(N3(g)). The total lattice potential energies are found to be: Epot(NaN3) = 725.1 kJ mole−1; Epot(KN3) = 650.7 kJ mole−1 and Epot(RbN3) = 632.1 kJ mole−1.  相似文献   

7.
讨论了NiCl2(bpy)3(bpy:2,2-联吡啶)在DMF中的电化学行为. 控制电位使电极过程处于扩散控制下, 采用计时电量法求得了29 ℃时NiCl2(bpy)3在DMF中的扩散系数为5.99×10-6 cm2•s-1, 不同温度下的扩散系数随温度升高而增大. 选择合适的电极电位, 使电极过程处于扩散和电化学混合控制下, 采用计时电量法求得了不同电极电位下的反应速率常数kf, 以及不同温度下的标准速率常数k0, 求得了表观活化能为14.4 kJ•mol-1.  相似文献   

8.
Thermal decomposition of poly(1,4-dioxan-2-one)   总被引:2,自引:0,他引:2  
To evaluate the feasibility of poly(1,4-dioxan-2-one) (PPDO) as a feed stock recycling material, the pyrolysis kinetics of PPDO were investigated. The pyrolysis of PPDO exclusively resulted in the distillation of 1,4-dioxan-2-one (PDO). From thermogravimetric measurements conducted at different heating rates, the kinetic parameters of the pyrolysis: activation energy, Ea=127 kJ mol−1; order of reaction, n=0; and pre-exponential factor, A=2.3×109 s−1, were estimated by plural analytical methods. The estimates show that the decomposition of PPDO proceeds by unzipping depolymerization as main reaction and random degradation process with lower Ea and A values. Equivalent isothermal degradation curves calculated from the thermogravimetric curves were supported by experimental isothermal degradation data. The calculation that PPDO is converted smoothly into PDO at 270°C agrees with the reported ceiling temperature of PPDO.  相似文献   

9.
Raman and infrared spectra of propylgermane, CH3CH2CH2GeH3, and its Ge-deuterated analog, CH3CH2CH2GeD3, were investigated in their gaseous, liquid and solid states. The normal coordinate treatment was carried out by density functional theory (DFT) calculation, using B3LYP/6-31G* and 6-311++G** basis sets, and the corresponding fundamental vibrations were assigned. The trans (T) and gauche (G) forms around the central C–C bond coexisted in the gaseous and liquid states and only the T form existed in the solid state. From the temperature dependent measurements of the Raman spectra in the liquid state, the enthalpy difference was found to be ΔH(TG)=−0.36±0.02 kcalmol−1 with the T form being more stable. The energy differences between the isomers obtained by DFT calculations were ΔE(TG)=−0.46 kcalmol−1 and ΔE(TG)=−0.87 kcalmol−1 by the 6-31G* basis set and 6-311++G** basis set, respectively.  相似文献   

10.
Tert-butyl chloride has been broadly studied experimentally through various techniques such as X-ray crystallography, DTA, and NMR. It was also studied experimentally through nuclear quadrupole resonance (NQR), but this study was limited and incomplete. In this paper, we present a more detailed study of TBC through the NQR of 35Cl. Our results show that near 120 K, the onset of the CH3 groups semirotations around symmetry axis C3 takes place with an activation energy U=16.1 kJ mol−1. This intramolecular movement produces a T1 minimum near 148 K and is the dominant mechanism of the nuclear spin-lattice relaxation in phase III of this compound. In phase II of TBC, we show that there are not only methyl groups semirotations, but also semirotations of the whole molecule around a different axis from the symmetry axis C3 (C–Cl bond) with an activation energy of E=10.4 kJ mol−1.  相似文献   

11.
The fracture toughness of impact modified polypropylene (PP) with impact modifier (IM) content up to 30 wt% was measured under quasi-static and impact rates of loading. Three types of IMs were employed in this work, which were all PP/PE based copolymers. Under quasi-static rate of loading, the fracture toughness for blends containing 30 wt% of IM were measured using the specific essential work of fracture (we) and the J-integral via the locus method (Jc). Both methods gave similar values of fracture toughness. Furthermore, the type of impact modifier used does not have a strong influence on the we (or Jc) value. The main difference between the three blends is the plastic work term (βwp). Using the specific essential work of fracture approach, no plane stress–plane strain transition can be observed. For impact rate of loading, the dynamic Gc and the Izod impact strength for blends containing 10–30 wt% of IM were measured. All the blends failed in a brittle or semi-brittle manner. In general, Gc increases with IM concentration.  相似文献   

12.
The five-coordinate mono-halide mononuclear Zn(II) complexes [Zn(tpa)X]+ (tpa = tris(2-pyridylmethyl)amine; X = I ([Zn(tpa)I]I; 1a), Br ([Zn(tpa)Br](ZnBr4)0.5; 2a) and Cl ([Zn(tpa)Cl](ZnCl4)0.5; 3a)) and the six-coordinate mononuclear complex [Zn(tpa)(NCS)2] (4a) have been synthesized and characterized by X-ray crystallography. The [Zn(tpa)X]+ complexes doped with the corresponding [Mn(tpa)X2] complexes (X = I (1b), Br (2b) and Cl (3b)) have been synthesized and their electronic properties investigated by multifrequency high field EPR (HF-EPR) (95–285 GHz). The magnetically diluted conditions allow the determination of the hyperfine coupling constant A (A = 68.10−4 cm−1 for 1b–3b). The zero-field splitting parameters (D and E) found for 1b–3b are comparable to those found for neat samples of the [Mn(tpa)X2] complexes (1b: D = 0.635 cm−1, E/D = 0.189; 2b: D = 0.360 cm−1, E/D = 0.192; 3b: D = 0.115 cm−1, E/D = 0.200). The efficacy of using multifrequency EPR under dilute conditions to precisely determine spin Hamiltonian parameters is discussed.  相似文献   

13.
Rotational-state distributions of the CO+ (A–X, B–X) and N2+(B–X) emissions produced by the collisions of He(2 3S) with CO and N2 were studied in the collision energy (ER range 100–200 meV. The rotational populations of the emitting states can be fitte by single Boltzmann temperatures (TR. The TR (320 ± 30 K) for the ν′ = 3 and 4 levels of the CO+ (A2Π) state are nearly independent of, or slightly increase with, ER, while TR for the CO+(B2Σ+, ν′ = 0) state increases rapidly with ER.The TR (430 ± 20 K) for the N2+(B2Σ+, ν′ = 0) state is nearly independent or slightly decreases with increasing ER. Interactions providing these trends are discussed.  相似文献   

14.
4-1,2:3,4-(trans-1,3,5-hexatriene)](η5-cyclopentadienyl)cobalt (3) undergoes dimerization to form a flyover carbene, 5, with concomitant elimination of one equivalent of trans-1,3,5-hexatriene. Structure 5 thermally rearranges via a metal-mediated [1,5]-H shift to carbene 6: Ea = 29.1 ± 0.4 kcal mol−1, log A = 11.6 ± 0.6. The structures of 5 and 6 were confirmed by single crystal X-ray determination. Low temperature irradiation of 6 generates 13 which undergoes a thermally induced reversion to 6: Ea = 19.4 ± 0.9 kcal mol−1, log A = 10.0 ± 1.3. Deuterium labeling studies indicate the mechanisms involved in these C---H transformations are intramolecular, regio-, and stereospecific. The chemical study of this system is extended to include a variety of homologous CpCo(triene) complexes. A comparison between the triene approach to the formation of these flyover pentadienyl carbenes and direct carbene addition is described.  相似文献   

15.
制备了吡啶类离子液体N-己基吡啶二氰胺盐[C_6py][DCA],并用核磁共振氢谱(~1H NMR)、核磁共振碳谱(~(13)C NMR)、差热扫描量热(DSC)、傅里叶变换红外(FT-IR)光谱对其进行表征。在288.15–338.15 K温度范围内,采用标准加入法,测定其密度(ρ)、表面张力(γ)和折光率(n_D)。在测得的实验数据的基础上,得到了离子液体[C_6py][DCA]的分子体积(V_m)、表面能(E_a)、摩尔极化度(R_m)和极化率(α_p)。结果显示E_a、R_m和α_p几乎不随温度的变化而发生改变。本文还提出了摩尔表面Gibbs自由能(g_s)的概念,并改进了E?tv?s方程。同时还计算了gs、临界温度(T_c)和E?tv?s方程经验参数(kE),并预测了离子液体[C_6py][DCA]的表面张力,预测值与实验值具有较好的一致性。  相似文献   

16.
The complex permittivity of a room-temperature ferroelectric liquid crystal 4-n-octyloxy benzoic acid 4'-[(2-methylbutyloxy)carbonyl]phenyl ester, has been measured in the vicinity of the phase transition in the frequency range 40 Hz-300 kHz. In the para-electric phase the contribution εs of the soft mode to the permittivity and the soft-mode relaxation frequency fs satisfy the Curie-Weiss law. Under a biasing field E, the helicoidal structure is unwound, and εs and fs can then be measured even in the ferroelectric phase. On the other hand, the phase transition is smeared under the influence of E. This smearing results in deviations from the Curie-Weiss law for both εs and fs in the vicinity of the transition. On increasing E, the maximum of the permittivity, εmax, is lowered and shifted to higher temperatures. Both the shift and ε-1max are proportional to E2/3. From experimentally found dependences, some constants in the free energy are determined.  相似文献   

17.
The temperature dependence of the rate constants, for the reactions of hydrated electrons with H atoms, OH radicals and H2O2 has been determined. The reaction with H atoms, studied in the temperature range 20–250°C gives k(20°C) = 2.4 × 1010M-1s1 and the activation energy EA = 14.0 kJ mol-1 (3.3 kcal mol-1). For reaction with OH radicals the corresponding values are, k(20°C) = 3.1 × 1010M-1s-1 and EA = 14.7 kJ mol-1 (3.5 kcal mol-1) determined in the temperature range 5–175°C. For reaction with H2O2 the values are, k(20°C) = 1.2 × 1010M-1s-1 and EA = 15.6 kJ mol-1 (3.7 kcal mol-1) measured from 5–150°C. Thus, the activation energy for all three fast reactions is close to that expected for diffusion controlled reactions. As phosphates were used as buffer system, the rate constant and activation energy for the reaction of hydrated electron with H2PO4- was determined to k(20°C) = 1.5 × 107M-1s-1 and EA = 7.4 kJ mol-1 (1.8 kcal mol-1) in the temperature range 20–200°C.  相似文献   

18.
The structure and kinetics of the crystallization reaction of amorphous Te51.3As45.7Cu3 were studied under nonisothermal conditions using scanning electron microscopy (SEM) and differential scanning calorimetry (DSC). Two exothermic changes were reported. Five isoconversional methods, of Kissinger–Akahira–Sunose (KAS), Flynn–Wall–Ozawa (FWO), Tang, Starink, and Vyazovkin, were used to determine the variation of the activation energy for crystallization with temperature, E(T). The results show that the activation energy for crystallization associated with the first peak first decreases with increasing temperature and then increases. Different behaviour was observed for the second peak, where an increase of E with temperature followed by a decrease. The effect of heating rate on the reaction model, g(), was also different for the two crystallization peaks.  相似文献   

19.
通过卷曲立方AlAs(111)单层片(sheets)构造了一系列(n,0)和(n,m)一维单壁纳米管。用周期性密度泛函理论(DFT)计算并比较了不同类型AlAs纳米管在几何结构、能量及电子性质等方面的差别。计算结果表明锯齿型和椅型纳米管应变能均为负值,并随着管径变大而逐渐变小。然而,它们的带隙相当不同:椅型纳米管为间接带隙,随着管径的增大而带隙减小;锯齿型纳米管为直接带隙,管径为1.87 nm时存在着一个极大带隙值(2.11 eV)。这种不同主要源于锯齿型纳米管铝原子间3p轨道的耦合贡献。  相似文献   

20.
Chakrabarti CL  Cathum SJ 《Talanta》1990,37(12):1111-1117
The mechanism of cobalt atomization from different atomizer surfaces in graphite-furnace atomic-absorption spectrometry has been investigated. The atomizer surfaces were pyrolytically coated graphite, uncoated electrographite, and glassy carbon. The activation energy of the rate-determining step in the atomization of cobalt (taken as the nitrate in aqueous solution) in a commercial graphite furnace has been determined from a plot of log ks vs. 1/T (for T values greater than the appearance temperature), where ks is a first-order rate constant for atom release, and T is the absolute temperature. The activation energy Ea, can be correlated either with the dissociation energy of CoO(g) or with the heat of sublimation of Co(s), formed by carbon reduction of CoO(s), the latter being the product of the thermal decomposition of Co(NO3)2. The mechanism for Co atomization seems to be the same for the pyrolytically coated graphite and the uncoated electrographite surfaces, but different for the glassy carbon surface. The suggested mechanisms are consistent with the chemical reactivity of the three atomizer surfaces, and the physical and thermodynamic properties of cobalt and its chemical compounds in the temperature range involved in the charring and atomization cycle of the graphite furnace.  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