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 78 毫秒
1.
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   

2.
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   

3.
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   

4.
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   

5.
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   

6.
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   

7.
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   

8.
The data on the permittivities of crystalline 1,4-dithiane and 1,4-dithiane-1,4-dioxide were used to calculate the molar deformation polarizations of the C-S (P ∞, C-S = 3.84 cm3/mol) and S$ \underline \ldots $ \underline \ldots O ($ P_{\infty ,S\underline \ldots O} $ P_{\infty ,S\underline \ldots O} = 4.34 cm3/mol) bonds within the framework of the additive scheme suggested by Levin. These data were used to calculate the deformation permittivity of dimethylsulfoxide (DMSO) at 298.15 K, ɛ∞, DMSO = 2.36, and the dipole correlation factor of pure DMSO, g dip = 1.055.  相似文献   

9.
Photooxidation of tryptophan (Trp) in complexes with the uranyl ion upon selective excitation was studied, and the quantum yields of amino acid oxidation (ϕ(O2)) were determined. It was shown that the photosensitized oxidation of Trp by the uranyl ion involves the chain reaction of Trp·+ radical cations with O2, rather than follows the commonly accepted mechanism of interaction of the substrate radical cation with the superoxide ion $ O\overline {_2^ - } $ O\overline {_2^ - } .  相似文献   

10.
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   

11.
Indirect methods of investigation of composition and defect structures of lithium niobate (LiNbO3) single crystals with different compositions are discussed. The analysis of two methods for the determination of the Li/Nb ratio in the samples is carried out, viz., the fundamental UV absorption edge and IR vibrational spectra of the OH group defects. Intrinsic defect concentrations in lithium niobate crystals (lithium vacancies, $ V_{Li^ - } $ V_{Li^ - } and defects, $ Nb_{Li^{4 + } } $ Nb_{Li^{4 + } } ) as a function of the Li/Nb ratio in the samples are given. The results obtained can serve as an effective way of express non-destructive composition analysis in a mass production of parallel-plane plates.  相似文献   

12.
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   

13.
Three cyanide-bridged dodecanuclear macrocyclic wheel-like complexes [Cr(bpmb)(CN)2]6[Mn(5-Brsalpn)]6·12H2O (1), [Co(bpmb)(CN)2]6[Mn(5-Brsalpn)]6·12H2O (2) and [Co(bpmb)(CN)2]6[Mn(5-Clsalpn)]6·24H2O·8CH3CN (3) [bpmb2−= 1,2-bis(pyridine-2-carboxamido)-4-methylbenzenate dianion; 5-Brsalpn2− = N,N′-propylenebis(5-bromosalicylideneaminato) dianion; 5-Clsalpn2− = N,N′-propylenebis(5-chlorosalicylideneaminato) dianion] have been synthesized and their crystal structures and magnetic properties have been investigated. The three compounds are structurally isomorphous and consist of alternating Mn(III)-Schiff base cations and [M(bpmb)(CN)2] anions, generating cyanide-bridged nanosized dodecanuclear macrocyclic structures with an approximate diameter of 2 nm. The study of the magnetic properties of complex 1 reveals an antiferromagnetic interaction between the Cr(III) and Mn(III) ions through the cyanide bridges. A best-fit to the magnetic susceptibility of the complex leads to a magnetic coupling constant of J CrMn = −2.65(6) cm−1 on the basis of a one-dimensional alternating chain model with the Hamiltonian $ H = - J_{CrMn} \sum\limits_{i = 0}^N {S_i \cdot S_{i + 1} } $ H = - J_{CrMn} \sum\limits_{i = 0}^N {S_i \cdot S_{i + 1} } .  相似文献   

14.
The content of oxygen in Ca0.6 − y Sr0.4La y MnO3 − δ, where y = 0 and 0.05, was determined by coulometric titration over the temperature range 650–950°C at oxygen partial pressure in the gas phase varied from 10−4 to 1 atm. The results were used to calculate the partial molar enthalpy, Δ$ \bar H $ \bar H O(δ), and entropy, Δ$ \bar S $ \bar S O(δ), of oxygen in manganites. Changes in the Δ$ \bar H $ \bar H O(δ) and Δ$ \bar S $ \bar S O(δ) dependences caused by the introduction of lanthanum are evidence of the formation of local clusters of the double perovskite type in the Ca0.6Sr0.4MnO3 − δ matrix.  相似文献   

15.
The title complexes (NH4)3[YIII(Nta)2] (I) (H3Nta = nitrilotriacetic acid) and {K[YIII(Egta)] · 4H2O} n (II) (H4Egta = ethyleneglycol-bis-(2-aminoethylether)-N,N,N′,N′-tetraacetic acid) were prepared, and their molecular and crystal structures were determined by single-crystal X-ray diffraction techniques. Complex I crystallizes in the rhombohedral crystal system with R $ \bar 3 $ \bar 3 c space group. The central Y3+ ion is eight-coordinated by two nitrogen and six oxygen atoms, which come from two tetradentate Nta ligands. The crystal data are as follows: a = 7.9340(14) ?, c = 54.611(15) ?, V = 2977.1(11) ?3, Z = 6, ρcalcd = 1.738 mg/cm3, μ = 3.011 mm−1, F(000) = 1596, R = 0.0234 and wR = 0.0641 for 686 observed reflections with I ≥ 2σ(I). The {K[YIII(Egta)] · 4H2O} n is nine-coordinated by two nitrogen and seven oxygen atoms and produces a 1D unlimited zigzag-type chain through a bridging carboxylic group. {K[YIII(Egta)] · 4H2O} n crystallizes in the monoclinic crystal system with C2/c space group. The crystal data are as follows: a = 37.588(5) ?, b = 13.7101(19) ?, c = 8.6070(12) ?, β = 99.929(2)°, V = 4369.0(11) ?3, Z = 8, ρcalcd = 1.753 mg/cm3, μ = 2.934 mm−1, F(000) = 2368, R = 0.0385 and wR = 0.0800 for 4082 observed reflections with I ≥ 2σ(I).  相似文献   

16.
The substitution of bis(2,4,6-tripyridyl 1,3,5-triazine)iron(II), \textFe(TPTZ) 2 2 + {\text{Fe(TPTZ)}}_{ 2}^{{ 2 { + }}} by 2,2′,6,2″-terpyridine (terpy) occurs on a time scale of about 6 m. The kinetics of this reaction was followed by stopped-flow spectrophotometry in the pH range of 3.6–5.6 in acetate buffer. The data indicate that the reaction occurs in two consecutive steps: kinetic data for both steps were acquired simultaneously and analyzed independently. The first step is assigned to the reaction between \textFe(TPTZ) 2 2 + {\text{Fe(TPTZ)}}_{ 2}^{{ 2 { + }}} and terpy to give Fe(TPTZ)(terpy)2+, followed by its reaction with another terpy molecule to give the final product, \textFe(terpy) 2 2 + {\text{Fe(terpy)}}_{ 2}^{{ 2 { + }}} . The rate of the reaction increases with increases in [terpy] and pH. The kinetic and activation parameters determined for both steps suggest that they involve both associative and dissociative paths. The ternary complex Fe(TPTZ)(terpy)2+ has been prepared, and the kinetics of its reaction with terpy suggest that this reaction is identical with the second step of the \textFe(TPTZ) 2 2 + {\text{Fe(TPTZ)}}_{ 2}^{{ 2 { + }}} -terpy system, supporting the proposed mechanism.  相似文献   

17.
Bis(2,4,6-tripyridyl 1,3,5-triazine)iron(II), \textFe(\textTPTZ) 2 2 + {\text{Fe(\text{TPTZ})}}_{ 2}^{{ 2 { + }}} reacts with 3-(2-pyridyl)-5,6-bis(4-phenyl-sulfonicacid)-1,2,4-triazine (PDTS) and 3-(4-(4-phenylsulfonicacid)-2-pyridyl)-5,6-bis(4-phenylsulfonic-acid)-1,2,4-triazine (PPDTS) to give \textFe(PDTS) 3 4- {\text{Fe(PDTS)}}_{ 3}^{ 4- } and \textFe(PPDTS) 3 7- {\text{Fe(PPDTS)}}_{ 3}^{ 7- } respectively. Both of these substitution reactions are fast and their kinetics were monitored by stopped-flow spectrophotometry in acetate buffers in the pH range of 3.6–5.6 at 25–45 °C. Both reactions are first order in \textFe(TPTZ) 2 2 + {\text{Fe(TPTZ)}}_{ 2}^{{ 2 { + }}} and triazine, and pH has negligible effect on the rate. The kinetic data suggest that these reactions occur in an associative path and a mechanism is proposed considering both protonated and unprotonated forms of PDTS and PPDTS are very similar in reactivity. The kinetic and activation parameters have been evaluated.  相似文献   

18.
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   

19.
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   

20.
The rapid oxidation ofbis(2,4,6-tripyridyl-1,3,5-triazine)-iron(II), [Fe(TPTZ)2]2+, bytrans-1,2-diaminocyclohexanetetraacetatomanganate(III), [MnIII(Y)], in acetate buffers was monitored using stopped-flow spectrophotometry. The reaction is first order in the substrate and evidence was obtained for pre-complexation between the oxidant and the substrate. The reaction rate increases as the pH increases. Characterisation of the products using the radiotracers54Mn and59Fe indicated that [MnII(Y)]2− and [Fe(TPTZ)2]3+ are the final products. The reaction obeys the rate law:
  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