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.
Optical methods of investigation of the defect structures and the composition of lithium niobate (LiNbO3) single crystalsis discussed. The intrinsic defects concentrations in lithium niobate crystals (lithium vacancies (VLi) and (NbLi)4+ defects, Nb on Li site in the valence state 4+), as a function of the Li/Nb ratio, are also reported. The optical damage resistance of various lithium niobate samples was investigated as a function of the composition. A remarkable increase in the optical damage resistance was found in MgO-doped almost stoichiometric lithium niobate crystals.  相似文献   

2.
The heat capacity and density of solutions of sodium and potassium perchlorates in N-methylpyrrolidone (MP) at 298.15 K were studied by calorimetry and densimetry. The standard partial molar heat capacities $ \bar C_{p2}^ \circ $ \bar C_{p2}^ \circ and volumes $ \bar V_2^ \circ $ \bar V_2^ \circ of NaClO4 and KClO4 in MP were calculated. The standard heat capacities $ \bar C_{pi}^ \circ $ \bar C_{pi}^ \circ and volumes $ \bar V_i^ \circ $ \bar V_i^ \circ of the perchlorate ion in an MP solution at 298.15 K were determined. The results are discussed with allowance for the specifics of solvation in the solutions of the salts under study. The coordination number of the ClO4 ion in an MP solution at 298.15 K was calculated.  相似文献   

3.
The transport properties of Sr5.66 Cu0.14Nb2.20 O11.30 double perovskite, which enters the homogeneity region of (Sr1-y Cu y )62x Nb2+2x O11+3x solid solution, are concerned. The total conductivity is differentiated into terms over wide ranges of temperatures and oxygen partial pressures $ p_{O_2 } $ p_{O_2 } in dry and humid atmospheres. When $ p_{O_2 } $ p_{O_2 } is low or high, a test sample has dominant electron transport of n- or p-type, respectively. In air ($ p_{O_2 } $ p_{O_2 } = 0.21 atm), the p-type electron conductivity term increases with temperature elevation. In a humid atmosphere ($ p_{H_2 O} $ p_{H_2 O} = 0.02 atm), a sample is capable of a reversible incorporation of water occlusion from the gas phase; as a result, some proton conductivity term appears and ion transference numbers increase over a wide range of $ p_{O_2 } $ p_{O_2 } values.  相似文献   

4.
A new approximation has been proposed for calculation of the general temperature integral $ \int\limits_0^T {T^m } e^{ - E/RT} dT $ \int\limits_0^T {T^m } e^{ - E/RT} dT , which frequently occurs in the nonisothermal kinetic analysis with the dependence of the frequency factor on the temperature (A=A 0 T m). It is in the following form:
$ \int\limits_0^T {T^m } e^{ - E/RT} dT = \frac{{RT^{m + 2} }} {E}e^{ - E/RT} \frac{{0.99954E + (0.044967m + 0.58058)RT}} {{E + (0.94057m + 2.5400)RT}} $ \int\limits_0^T {T^m } e^{ - E/RT} dT = \frac{{RT^{m + 2} }} {E}e^{ - E/RT} \frac{{0.99954E + (0.044967m + 0.58058)RT}} {{E + (0.94057m + 2.5400)RT}}   相似文献   

5.
A rapid, relatively sensitive and simple kinetic-spectrophotometric method for determining sodium salicylate content has been developed and validated. This method was based on the Fenton reaction that involved mixing of ferrous ions, hydrogen peroxide and sodium salicylate in acetic buffer medium. Ferrous ions, oxidized by hydrogen peroxide, formed with salicylate anion purple complex whose degradation started immediately. This effect was monitored by the decrease of absorbance at 525 nm. The experimental results showed that the most favorable conditions for complex degradation are: pH = 3.9, $ c_{Fe^{2 + } } $ c_{Fe^{2 + } } = 4.7 × 10−5 M, and $ c_{Fe^{2 + } } /c_{H_2 O_2 } $ c_{Fe^{2 + } } /c_{H_2 O_2 } = 1/150. The activation energy (E a) calculated from the slope ($ - \frac{{E_a }} {{2.303R}} $ - \frac{{E_a }} {{2.303R}} = 2.84 × 103) was 54.3 ± 0.6 kJ/mol. The absorbance increased linearly with the increment of sodium salicylate concentration (r = 0.9983). The system obeyed Beer’s law in a range of 0.93–9.3 μg/mL of sodium salicylate concentration. The calculated values for the detection limit, according to two formulas, available in the literature, were found to be 0.67 and 0.48 μg/mL. The variables affecting the rate of the proposed reaction were investigated. The relative standard deviations for five-replicate determinations of 0.93, 3.31, and 9.30 μg/mL of sodium salicylate were calculated to be 6.80, 2.95, and 1.71%, respectively. The proposed method has been successfully applied to determining sodium salicylate in human serum and wine and validated by HPLC (high-pressure liquid chromatography) reference method.  相似文献   

6.
Potentiometric method was used to measure the redox potentials of Tm3+/Tm2+ in a eutectic melt of sodium, potassium, and cesium chlorides relative to a chlorine reference electrode in the temperature range 823–973 K. The main thermodynamic characteristics of the redox reaction TmCl2(solution) + 1/2Cl2(g) ⇆ TmCl3(solution) were calculate from the conditional standard potentials $ E_{{{Tm^{3 + } } \mathord{\left/ {\vphantom {{Tm^{3 + } } {Tm^{2 + } }}} \right. \kern-\nulldelimiterspace} {Tm^{2 + } }}}^* $ E_{{{Tm^{3 + } } \mathord{\left/ {\vphantom {{Tm^{3 + } } {Tm^{2 + } }}} \right. \kern-\nulldelimiterspace} {Tm^{2 + } }}}^* .  相似文献   

7.
The generalized temperature integral $ \int\limits_0^T {T^m } \exp ( - E/RT)dT $ \int\limits_0^T {T^m } \exp ( - E/RT)dT frequently occurs in non-isothermal kinetic analysis. Here E is the activation energy, R the universal gas constant and T the absolute temperature. The exponent m arises from the temperature dependence of the pre-exponential factor. This paper has proposed two new approximate formulae for the generalized temperature integral, which are in the following forms:
$ \begin{gathered} h_m (x) = \frac{x} {{(1.00141 + 0.00060m)x + (1.89376 + 0.95276m)}} \hfill \\ h_m (x) = \frac{{x + (0.74981 - 0.06396m)}} {{(1.00017 + 0.00013m)x + (2.73166 + 0.92246m)}} \hfill \\ \end{gathered} $ \begin{gathered} h_m (x) = \frac{x} {{(1.00141 + 0.00060m)x + (1.89376 + 0.95276m)}} \hfill \\ h_m (x) = \frac{{x + (0.74981 - 0.06396m)}} {{(1.00017 + 0.00013m)x + (2.73166 + 0.92246m)}} \hfill \\ \end{gathered}   相似文献   

8.
The volta potential difference method at 298.15 K was used to determine the real primary medium effect for magnesium, calcium, cadmium, and copper ions, and also the real Gibbs transfer energy of these ions from water into a mixed water ethanol (EtOH) solution. The surface potential value at the nonaqueous solution/gas phase interface $ \Delta \chi _{H_2 O}^{EtOH} $ \Delta \chi _{H_2 O}^{EtOH} was obtained. With account for this value, chemical thermodynamic characteristics of the studied ions in the water-ethanol solvent were calculated and the effect of composition and nature of the mixed solvent on the values obtained was analyzed. The dependence of variation in the thermodynamic characteristics of cation resolvation was established on their crystallographic radius that corresponds to the following sequence: Ca2+ < Cd2+ < Cu2+ < Mg2+.  相似文献   

9.
Theoretical study of several O-nitrosyl carboxylate compounds have been performed using quantum computational ab initio RHF and density functional B3LYP and B3PW91 methods with 6-31G** basis set. Geometries obtained from DFT calculations were used to perform the natural bond orbital (NBO) analysis. It is noted that weakness in the O3-N2 bond is due to $ n_{O_1 } \to \sigma _{O_3 - N_2 }^* $ n_{O_1 } \to \sigma _{O_3 - N_2 }^* delocalization and is responsible for the longer O3-N2 bond lengths in O-nitrosyl carboxylate compounds. It is also noted that decreased occupancy of the localized $ \sigma _{O_3 - N_2 } $ \sigma _{O_3 - N_2 } orbital in the idealized Lewis structure, or increased occupancy of $ \sigma _{O_3 - N_2 }^* $ \sigma _{O_3 - N_2 }^* of the non-Lewis orbital, and their subsequent impact on molecular stability and geometry (bond lengths) are related with the resulting p character of the corresponding sulfur natural hybrid orbital (NHO) of $ \sigma _{O_3 - N_2 } $ \sigma _{O_3 - N_2 } bond orbital. In addition, the charge transfer energy decreases with the increase of the Hammett constants of subsitutent groups.  相似文献   

10.
The kinetics of methoxy-NNO-azoxymethane (I) hydrolysis in concentrated solutions of strong acids (HBr, HCl, HClO4, and H2SO4) has been investigated by a manometric method. The gas evolution rate is described by the equation corresponding to two consecutive first-order reactions, with the rate constant of the second reaction considerably exceeding the rate constant of the first reaction, i.e., k 2 {ie17-1} k 1. The temperature dependences of k 1 (s−1) in 47.59% HBr in the temperature range from 60 to 90°C and in 64.16% H2SO4 between 80 and 130°C are described by Arrhenius equations with IogA= 12.7 ± 1.5 and 13.6 ± 1.4 and E a = 115 ± 10 and 137 ± 10 kJ/mol, respectively. The parameters of the Arrhenius equation for the rate constant k 2 for the reaction in 64.16% H2SO4 between 80 and 130°C are IogA= 9.1 ± 2.5 and E a = 91 ± 18 kJ/mol. An analysis of the UV spectra of compound I in concentrated H2SO4 shows that I is a weak base $ (pK_{BH^ + } \approx - 6) $ (pK_{BH^ + } \approx - 6) . The rate-determining step of the hydrolysis of I is the attack of the nucleophile on the carbon atom of the MeO group of the protonated molecule of I. The resulting methyldiazene dioxide decomposes via a complicated mechanism to evolve N2, NO, and N2O. The pseudo-first-order rate constant k 1 of the reaction at 80°C depends strongly on the acid concentration and on the type of nucleophile (Br, Cl, or H2O). The relationship between k 1 and the rate constant k of the bimolecular nucleophilic substitution reaction (SN2) is given by the linear equation log$ [k_1 /(C_H + C_{Nu} )] = m^ \ne m*X_0 + \log (k/K_{BH^ + } ) $ [k_1 /(C_H + C_{Nu} )] = m^ \ne m*X_0 + \log (k/K_{BH^ + } ) , where $ C_{H^ + } $ C_{H^ + } and C Nu are the concentrations of H+ and nucleophile, respectively; X 0 is the excess acidity; and m and m* are coefficients. The Swain-Scott equation log$ (k_{Nu} /k_{H_2 O} ) = ns $ (k_{Nu} /k_{H_2 O} ) = ns , where n is the nucleophilicity factor and s is the substrate constant (s = 0.72), is applicable to the rate constants k of the SN2 reactions of the protonated molecule of I with Br, Cl, and H2O.  相似文献   

11.
The phase composition and microstructure of La1 − x Ca x MnO3 (x = 0–1) materials prepared by the Pechini method from polymer-salt stocks were studied after testing these materials in methane oxidation. According to X-ray diffraction data, the reaction medium causes no significant changes in the samples, while high-resolution transmission electron microscopy indicates that the x > 0.3 samples are unstable. Under the action of the reaction medium, the perovskite structure of these samples undergoes partial decomposition accompanied by the formation of planar defects having a lower manganese content. The number and degree of segregation of these defects increase with increasing calcium content. The calcium oxide and manganese oxide phases as segregated nanoparticles are observed on the particle surface. These changes are caused by the decrease in the oxygen content of the manganites under the action of the reaction medium $ (T,P_{O_2 } ) $ (T,P_{O_2 } ) , by the formation of vacancies, and by the variation of the charge of the manganese cations, as well as by the charge ordering tendency of the manganese cations. Therefore, the observed changes in catalytic activity under the action of the reaction medium for x > 0.3 can be due to perovskite decomposition accompanied by the formation of planar defects, the release of the manganese oxide and calcium oxide phases, and their subsequent sintering.  相似文献   

12.
The adsorption of Cl, Br, and I ions on the renewable liquid In-Ga and Tl-Ga electrodes from 0.1 M solutions in dimethyl formamide (DMF) is investigated by using the method of differential capacitance measurements. The results are compared with similar data obtained on Hg and Ga electrodes in DMF and with the corresponding data obtained in acetonitrile (AN). It is shown that, in DMF, the adsorption parameters and the series of surface activity of halide ions (Hal) significantly depend on the metal nature. In contrast to Hg electrode, on which the surface activity of halide ions increases in the series: Cl < Br < I, on In-Ga, as well as on the Ga electrode, it varies in the reverse order: I < Br < Cl, whereas on the Tl-Ga electrode, partially reversed series of surface activity is observed: Br < I < Cl. The results are explained within the framework of Andersen-Bockris model. An analysis of experimental results leads to the following qualitative conclusions: (1) on the In-Ga and Tl-Ga electrodes, as well as on Ga electrode, free energy of metal-Hal interaction ( $ \Delta G_{_{M - Hal^ - } } $ \Delta G_{_{M - Hal^ - } } ) increases in series I < Br < Cl; (2) for Cl, Br, and I, $ \Delta G_{_{M - Hal^ - } } $ \Delta G_{_{M - Hal^ - } } ) grows in series Tl-Ga < In-Ga < Ga; (3) an absolute magnitude of $ \Delta G_{_{M - Hal_1^ - } } - \Delta G_{_{M - Hal_2^ - } } $ \Delta G_{_{M - Hal_1^ - } } - \Delta G_{_{M - Hal_2^ - } } (Hal1, and Hal2 are any ions of Cl, Br, and I) increases in series Hg < Tl-Ga < In-Ga < Ga; (4) the metal-DMF chemisorption interaction is much stronger than the metal-AN interaction and increases in series Tl-Ga < In-Ga < Ga.  相似文献   

13.
A method for estimating the critical temperatures (T b) of thermal explosion for energetic materials is derived from Semenov’s thermal explosion theory and the non-isothermal kinetic equation dα/dt=A 0 T B f(α)e−E/RT using reasonable hypotheses. The final formula of calculating the value of T b is $ \left( {\frac{B} {{T_b }} + \frac{E} {{RT_b^2 }}} \right) $ \left( {\frac{B} {{T_b }} + \frac{E} {{RT_b^2 }}} \right) (T bT e0=1. The data needed for the method, E and T e0, can be obtained from analyses of the non-isothermal DSC curves. When B=0.5 the critical temperature (T b) of thermal explosion of azido-acetic-acid-2-(2-azido-acetoxy)-ethylester (EGBAA) is determined as 475.65 K.  相似文献   

14.
Heat capacities of solutions of alkali metal iodides (MeI) in N-methylpyrrolidone (MP)-water mixed solvent were measured over the range of compositions. The influence of the composition of the mixed solvent on the heat capacity of MeI-MP-H2O ternary systems is discussed. Standard partial molar heat capacities $ \bar C_{p_2 }^o $ \bar C_{p_2 }^o (MeI) in the MP-water mixed solvent at 298.15 K are calculated.  相似文献   

15.
Processes occurring in Nb6O198−-WO42−-H+-H2O system where c Nb: c W = 4: 2, c Nb+W0 = 5 × 10−3, 2.5 × 10−3, or 10−3 mol/L, and ionic strengths I = 0.01–0.14 are created by NaCl background electrolyte were studied by pH titration and mathematical modeling. Solute ion species distribution diagrams were obtained for $ Z = \frac{{c_{H^ + }^0 }} {{c_{Nb + W}^0 }} = 0 - 1.5 $ Z = \frac{{c_{H^ + }^0 }} {{c_{Nb + W}^0 }} = 0 - 1.5 . The concentration constants and thermodynamic constants of formation were calculated for isopolyniobotungstate anions (IPNTAs). H x Nb4W2O19(6−x)−, (x = 1–5), ions were shown to appear in solution only after Nb6O198− was protonated and aquapolytungstate anions were formed. The results of modeling were supported by the synthesis of Tl3H3Nb4W2O19 · 16.5H2O, Tl2H4Nb4W2O19 · 11H2O, and NaTl3(H4Nb4W2O19)2 · 22H2O salts, which were identified by chemical analysis and IR spectroscopy.  相似文献   

16.
The kinetics of acrylic acid oxidation in the presence of iron ions $ (T = 333K,_{P_{O_2 } } = 1 atm) $ (T = 333K,_{P_{O_2 } } = 1 atm) has been investigated by measuring the oxygen uptake. The reaction has an induction period τ, after which the O2 uptake is described by the parabolic kinetic law δ[O2]0.5 = b(t − τ). The parameter b characterizing the catalytic oxidation of acrylic acid has been calculated. Upon the introduction of an initiator (azobisisobutyronitrile), the oxidation has no induction period, but the autoacceleration of the reaction is still observed. A mechanism is suggested for the process. This mechanism includes initiation due to the interaction of the resulting peroxide and hydroperoxide groups with Fe2+ and Fe3+ ions and chain termination via the reaction R· + Fe3+, where R· is an acrylic acid macroradical.  相似文献   

17.
Density of the water-ethylene glycol binary mixtures was measured in the entire range of compositions in the temperature range 278–333.15 K (6 values) at atmospheric pressure using a vibration densimeter. Mixtures with low concentrations of ethylene glycol were studied at 15 temperatures in the range of 274–333.15 K. Excess molar volumes V m E , the partial molar volumes of water -V 1 and ethylene glycol, -V 2, the coefficients of thermal volume expansion α of the mixture, the partial molar volume coefficients of thermal expansion of water $ \bar V_1 $ \bar V_1 and ethylene $ \bar V_2 $ \bar V_2 were calculated. Excess molar volumes were described using the Redlich-Kister equation. The density ρ of the mixture was found to increase with the increasing ethylene glycol concentration at all temperatures, but at low content of ethylene glycol the dependence ρ = f(T) of the mixture at ∼276.5 K passed through a maximum. The coefficient α increases sharply in the composition range 0 < x < 0.2, in the range 0.5 < x <1 remains almost unchanged, and at T > 277 K is positive for all compositions. The dependences $ \bar \alpha _1 $ \bar \alpha _1 = f (x) and $ \bar \alpha _2 $ \bar \alpha _2 = f (x) are complex in whole temperature range and are characterized by the presence of an extremum. V m E values are negative at all temperatures, and upon increase in the temperature the deviation from ideality decreases (x is the mole fraction of ethylene glycol).  相似文献   

18.
Densities, ρ 123, and speeds of sound, u 123, of 1-methyl pyrrolidin-2-one (1) + benzene or methyl benzene or cyclohexane (2) + propan-2-ol (3) ternary mixtures have been measured over the entire composition range at 308.15 K and atmospheric pressure. The resulting ρ 123 and V123EV_{123}^{\mathrm{E}} data were utilized to predict excess isentropic compressibilities, (kSE)123(\kappa_{S}^{\mathrm{E}})_{123}, of the studied (1+2+3) mixtures. The observed V123EV_{123}^{\mathrm{E}} and (kSE)123(\kappa_{S}^{\mathrm{E}})_{123} data have been analyzed in terms of Graph theory (which involved the topology of a molecule). It has been observed that V123EV_{123}^{\mathrm{E}} and (kSE)123(\kappa_{S}^{\mathrm{E}})_{123} values determined by Graph theory compare well with their corresponding experimental values.  相似文献   

19.
The lithium salt of the bis-furyl substituted disilazide anion, Li{i} [{i} = N(SiMe2R)2 where R = 2-methylfuryl] has been examined as a ligand transfer reagent for the synthesis of group 2 (magnesium) and group 13 (aluminium) compounds. Salt metathesis between Li{i} and AlMe2Cl afforded the expected dimethyl species, Al{i}Me2 (1), which was isolated as a colourless oil. In contrast the corresponding aluminium dichloride, synthesized from Li{i} and AlCl3, gave crystalline products as both the THF adduct Al{i}Cl2(THF) (2a) and the base-free derivative, Al{i}Cl2 (2b). The homoleptic magnesium bis(amide) Mg{i}2 (3) was also synthesized. X-ray crystallographic analysis of 2a reveals a four-coordinate distorted tetrahedral metal, in which neither of the furyl-substituents interact with the metal. In contrast, the aluminium in the base-free dichloride 2b is five-coordinate, containing the first structurally characterized example in which the amide binds with a κ1N,O,O′-bonding mode, involving coordination of both furyl-substituents at the N-bound metal.  相似文献   

20.
The applicability of the equation $ e^{ - ((r_1 - r_0 )/b)^{5/3} } + e^{ - ((r_2 - r_0 )/b)^{5/3} } = 1 $ e^{ - ((r_1 - r_0 )/b)^{5/3} } + e^{ - ((r_2 - r_0 )/b)^{5/3} } = 1 has been studied. The equation defines the relationship between the experimental values of the covalent (r 1) and hydrogen (r 2) bond lengths in O-H...O bridges for describing the relation between the experimental interatomic distances in N-H...N bridges and the parameters of X-H...X fragments (X = O, N, F, Cl) calculated by the density functional method (B3LYP/6-31++G(d,p)) for neutral, positive, and negative molecular complexes. Here r 0 is the mean value of the X-H bond length in free molecules; r sym is the X...H distance in the symmetrical bridge; and b is the coefficient defined by the equation b = (r symr 0)/(ln2)3/5. This equation allows us to adequately describe the relationships between bond lengths in nearly linear hydrogen bridges formed by oxygen, nitrogen, fluorine, and chlorine atoms. It is thus universal and can be used in studies of a wide range of substances.  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