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.
The room-temperature diffuse-reflectance spectra of compositions within the Li1+xTi2?xO4 spinel system (0 ≤ x ≤ 13) show three absorption bands in the range 4000 to 48,000 cm?1. Two high-energy absorption bands correspond to charge-transfer transitions from the oxygen-2p valence band to the titanium t2g and σ1 conduction bands, where the σ1 band of eg character has hybridized titanium-3d and titanium-4s parentage. The absorption band arising from promotion of electrons to the empty σ1 band does not alter with composition whereas the absorption band arising from promotion of electrons to the partially filled t2g band narrows as the concentration of conduction electrons in the t2g band decreases. These two high-energy absorption bands fall entirely within the ultraviolet spectral region, and the absorption edge in Li43Ti53O4 (x = 13) occurs at 24,300 cm?1 (3.02 eV). A low-energy absorption band is observed in compositions with x < 13 and in samples of Li43Ti53O4 reduced in hydrogen at elevated temperatures. This band straddles the boundary between the visible and infrared spectral regions and shifts toward lower energy as the concentration of conduction electrons in the t2g band decreases. The possible origins of the band are discussed; the argument is in favor of a d-d interband transition from states in the partially filled t2g band to states in the empty σ1 band.  相似文献   

2.
By monitoring the absorbance of a solution in which an absorbing species is undergoing first- or pseudo-first-order decay, one can evaluate both the initial absorbance A0 and the rate constant k. The precisions of the resulting values depend on the standard errors of measurement of both absorbance and time, which are most conveniently expressed by the dimensionless parameters σA/A0 and kσt. Equations describing these dependences for both the overall precision indexes for the individual parameters and the global overall precision index are given to aid in maximizing the efficiencies and precisions of practical experiments.  相似文献   

3.
A mathematical analysis is developed to take into account the motion of the electrolyte/electrode boundary during electrolysis at constant current or at constant potential. The change in volume of the electrode is deduced from the partial molal volume of the electroactive component, which is supposed to be constant during the experiment.The applicability of this treatment is discussed. It is shown that, after a potential step, the decrease of the current density is a linear function of t?12, as for the Cottrell law, but the proportionality factor is changed. During the passage of a constant current the surface concentration remains nearly a linear function of t12, as for the Sand law when the concentration of the electroactive component is small. In general, very large departures from the standard equations appear at high concentrations of electroactive component. Applications are given for measuring solid state interdiffusion coefficients from electrochemical techniques. This study can be used to determine the kinetics of electrolyte diffusion coating and surface alloying.  相似文献   

4.
A beam of state-selected NO molecules (J = Ω = 32) has been produced by an electrostatic hexapole and has been collided with O3 molecules in a scattering chamber. The E-field dependence of the chemiluminescent cross section, σhr, has been investigated and resulted in the determination of the M-dependence of σhr: σhr (M)/σ0 = 1.192±0.009, 0.0848±0.015, 1.177±0.015, 0.783±0.009 for M = 32, 12, ?12 and ?32, respectively. Application of the Legendre expansion technique and the density matrix formalism provided a deconvoluted σhr(γ), for a single angle of attack γ of the NO axis, expressed in simple model functions with adjustable parameters. From this analysis it is concluded that chemiluminescence only occurs when cos γ ≈ 1, the “end-on-head” orientation of NO yielding ≈ 30% of all collected light, and when cos γ ≈ ?0.275, the “broad-side-tail” orientation of NO yielding the remaining 70%. The steric factors belonging to these reactive orientations have been estimated and are S1 = 0.25±0.07 and S2 = 0.40±0.09, respectively. The observed dependence of σhr has been confronted with the rules of Woodward and Hoffman. Although there are indeed two symmetries (bpl and cpl) correlating the electron orbitals of the reactants and the products, these rules do not lead to an explanation of the steric effects of the NO+O3 reaction.  相似文献   

5.
Calibration methods differ as to the number and concentration of calibration standards, and whether these are added to the samples or separate from them. The four main calibration mehods (single separate or added standard and multiple separate or added standards) and some modifications are desribed mathematically and subjected to error-propagation analysis, to examine the likely effects of errors in the analytical signal on the overall accuracy and precision of the concentration estimate. Comparison of the results throws light on the influence of the number, concentration and nature of the calibration standards, the effects of sample and standard replication, and the costs and the benefits of blank measurements. It is shown that all standard-addition methods are immune to proportional signal error, but more sensitive to nonlinearity. In separate-standard methods, all bias disappears when the true sample concentration (xs) is equal to the standard concentration or to the mean standard concentration (x). Only the multiple separate standard method is unaffected by constant error common to sample and standards, without blanking. In multiple-standard methods, precision is best at xs = x. Precision is always improved by increasing the number of sample and standard measurements; standard-addition methods respond best to sample replication. Whichever calibration method is used, recovery correction will eliminate proportional concentration error, at the cost of decreased precision.  相似文献   

6.
7.
Various types of photometric titration curves are discussed. If a metal M is titrated conipleximetrically using a metal indicator and the absorbance is plotted vs. the titrant consumed, the inflection point appears at a pM value defined by the equation 3 pMinfl = pMtrans + 2 pMeqThis expression is valid when M combines in a 1 : 1 ratio with the complexing agent and the indicator and when the indicator concentration is small compared to the total metal concentration.The difference between the pM values at the inflection and equivalence points can be calculated from the equation ΔpM = pMinfl — pMeq = 13(pMtrans — pMeq) = 16log(CMK2MI/KMY)If the inflection point is taken as the equivalence point, the error arising can be calculated from ΔpM, or more simply, read from a diagram.If transmittance, instead of absorbancc, is plotted as a function of the titrant volume, the inflection point depends on the added amount of indicator. However, at high transmittance values, i.e., at low indicator concentrations, the inflection point of a transmittance curve occurs practically at the same volume of added titrant as the inflection point of an absorbance curve. Rules are given for applying an indicator correction for the amount of metal bound to the indicator at the end-point.The derived equations and discussions can also be applied to acid-base titrations.  相似文献   

8.
Separate optimized s and p bond functions (BFs) were added to the corresponding Dunning basis sets (11s,6p15s) → (5s,4p13s) for the four hydrides. Properties calculated with these basis sets are quasi-identical to those obtained with conventional polarization functions (dlp). The computer time ratios found are: t(BF)/t(dlp) = 12 for UHF calculations, and 23 for MC SCF calculations.  相似文献   

9.
The title compound is the final product of the reaction between dimethylacetylene dicarboxylate and [(η5-C5H5)W(SMe) (CO)3 ]. Its molecular structure has been established by an x-ray analysis based on 3301 diffractometric intensities. The crystals are monoclinic, space group P21/a, with four molecules in a cell of dimensions a = 10.323(2), b = 16.016(2), c = 10.437(2)», β = 103.56(2)°. After least-squares refinement R = 0.036. The tungsten co-ordination is square-pyramidal, the apical site being occupied by the centroid of the cyclopentadienyl ring. The basal co-ordination sites contain two carbonyl groups and also the sulphur and σ-bonded carbon atoms of a chelating carbothiolic methyl ester ligand derived from the incoming alkyne and CO and SMe groups of the original complex. The W-S and σ-W-C bond lengths are 2.440(2) and 2.194(7)».  相似文献   

10.
The rates of solvolysis of 2 - aryl - exo - 5,6 - trimethylene - exo - and endo - 2 - norbornyl p-nitrobenzoates (7 and 8, respectively) with representative substituents [p-CH3O, p-CH3, H, m-CF3, p-CF3, and 3,5-(CF3)2] were determined in 80% aqueous acetone and compared with those of the parent 2-aryl-exo- and endo-2-norbornyl p-nitrobenzoates (5 and 6, respectively). The rate ratios for the endo-p-nitrobenzoates (68) are essentially constant and close to unity for these substituents, indicating that the perturbation of the trimethylene bridge toward the C2-position is virtually negligilbe. The rate ratios for the exo-p-nitrobenzoates (57) can also be regarded as invariant over the reactivity range studied. The exo/endo rate ratios (78) are 246 (p-CH3O), 196 (P-CH3), 129 (H), 80 (m-CF3), 90 (p-CF3), and 89 [3,5-(CF3)2], being similar to the corresponding 56 rate ratios. The solvolyses of these p-nitrobenzoates (7 and 8) afford predominantly ( > 97%) exo-alcohols. Since the secondary exo-5,6-trimethylene-2-norbornyl system, with its low exo/endo rate ratio, 11.2, is known to solvolyse without significant σ-participation, the tertiary derivatives should also undergo solvolysis without σ-participation. Consequently, the similarities in the solvolytic behaviors between the two systems (5 vs 7; 6 vs 8) strongly support the previous conclusion that σ-participation is unimportant in the solvolysis of 5.  相似文献   

11.
Measurements of dc conductivity of Cd-stearate multilayer films below room tenperature down to liquid nitrogen temperature have revealed the temperature dependence σ = σ0 exp(?AT?12). This relation, together with the values of σ0 and A, confirms the theoretical predictions for hopping conduction in a multilayer system associated with interface states.  相似文献   

12.
Dehydrotripeptides, X-C*CH(Ri)-ΔPhe-C*CH(Rk)COOMe (3: X = tBOC-NH,CBZ-NH, CF3CONH and N3), were employed for the asymmetric hydrogenation catalyzed by chiral rhodium complexes and it was found that tBOC-3 brought about by far the best results. Stereoselective dideuteration of a tBOC-3 was successfully performed.  相似文献   

13.
cis- and trans - 2,3 - Dimethylenemethylenecyclopropane (C and T) interconvert at 160.0° with a small normal kinetic isotope effect (KIE) when the exo-methylene is deuterated, but the 1,3-shift products, 2-methylethylidenecyclopropane, show a large normal KIE, 1.35 and 1.31, when formed from C and T, respectively. This data can be interpreted in terms of either parallel reactions or a common trimethylenemethane diradical intermediate formed with a normal KIE of 1.11 and closing to 1,3-shift product with a normal KIE of 1.29 due to the effect of deuterium in the required 90° rotation of the exo-methylene carbon.The kinetics of the thermal 1,3- and 3,3-shifts of cis- and rans-3,4-dimethyl-1,2-dimethylenecyclobutane (CB and TB) were determined in a flow reactor. The first order rate constants are log kCB (sec?1) = 13.7 ? 42,200/2.3 RT and log kTB (sec?1) = 13.6 ? 41,900/2.3 RT (Ea in kcal/m) which compare favorably to that from the parent hydrocarbon. 1,2-dimethylenecyclobutane, after reasonable correction for dimethyl substitution.Rearrangement of TB and its bis(dideuteriomethylene) derivative at 230.0° revealed a normal KIE of 1.08. This KIE could be interpreted in terms of either a methylene rotational isotope effect in a concerted reaction or formation of a bisallyl diradical with the expected normal rotational IE on closure to the 1,3-shift product of 1.12 with no IE in the ring opening when the result is corrected for return of the biradical to starting material.The kinetics of intramolecular 2 + 2 cycloaddition of 1,2,8,9-decatetraene were determined in a flow reactor. The first order rate constant is log k(sec?1) = 9.4 ? 30,800/2.3 RT (Ea in kcal/m). These energetics are compared with those of other 2 + 2 cycloadditions. The major product is 3,4-dimethylenecyclooctene (DC) which is also found from the minor product, cis-7,8-dimethylenebicvyclo[4.2.0]octane (CO), at higher temperatures. The trans isomer, TO, also gives DC at about the same rate as CO.  相似文献   

14.
The structure of wüstite Fe1?zO is studied by neutron diffraction on one polycrystalline sample under equilibrium conditions at high temperature. The mean isotrope temperature factor B is expressed as a sum of two parts, BTh and BSt, which vary linearly with a single parameter, respectively temperature and z. A classification is established for clusters (mn) settled from m vacancies in octahedral sites and n FeIII ions in interstitial sites. Sixteen values have been experimentally determined for the vacancies to interstitials ratio ? = (z + t)t = mn. A constant value of ?, which is lower than 3, is observed. This result characterizes the short-range order. It eliminates several possibilities of clusters like those obeying the relation ? = (1 + 3n)n. Other clusters, namely (166) or (4014), might agree. The (83) and (94) clusters obtained from (41) clusters joined by an edge would be the more likely. An analysis of diffuse scattering eliminates the hypothesis of large domains with inverse spinel structure. The structural differences between the three varieties W1, W2 and W3 would not be found in a structural change of clusters.  相似文献   

15.
Pentafluorobenzoyl peroxide (FBP) reacted rapidly with dimethoxybenzenes in F113 (CCl2FCClF2) with kinetics of first order in each component. High yields of ring-substituted esters of m-dimethoxybenzene (m-DMB) and p-dimethoxybenzene (p-DMB) were obtained, whereas for 2,5-di-t-butyl-1,4-dimethoxybenzene (DBDMB), the t-Bu group was simultaneously eliminated. For 2,5-dimethyl-1,4-dimethoxybenzene (DMDMB), the benzylic hydrogen was substituted. Rate and product studies both indicate a rate-determining electron transfer step leading to radical ion pairs which collapse to products.  相似文献   

16.
17.
The reaction between AuCl(PPh3 and [Fe(μ3-HCNBut)(CO)9]? gives AuFe33-HCNBut)(CO)9(PPh3), crystals of which are triclinic, space group P1, with a 12.815(3), b 16.265(4), c 19.106(3) Å, α 67.15(3), β 73.46(2), γ 73.12(2)° and Z = 4. The comlex contains an AuFe3 “butterfly” cluster, the Fe3 face of which is bridged by the HCNBut ligand bonded in the (2σ + π) mode on the side opposite to the Au(PPh3) moiety, which, contrary to expectation based on the analogy with H, bridges the two Fe atoms σ-bonded to N, and π-bonded to the CN group. The AuFe2/Fe3 dihedral angles in the two independent molecules differ significantly, with values of 110.9 and 132.1°.  相似文献   

18.
(C5H5)2NbBH4 reacts with C5H5M(CO)3Me in toluene solution in the presence of Et3N to give binuclear complexes (C5H5)2NbM(CO)3C5H5 where M is Mo or W (IV and V, respectively). The structure of IV has been studied by X-ray diffraction (the crystals are orthorhombic, a 12.748(5), b 16.745(6), c 14.314 A/ac>?;; Z = 8, space group of Pbca, automatic difractometer Syntex P2I, λ(Mo-Kα, 1382 reflections, R = 0.056, Rw = 0.058). Molecule IV contains a wedge-like sandwich (π-C5H5)2Nb (NbC 2.37–2.48, CC (av) 1.42 A/ac>?;, angle between ring planes 49°) linked with the (π-C5H5)Mo(CO) fragment by a direct NbMo bond (3.073 A/ac>?;) and two bridging CO groups, one nonsymmetrically bonded through the carbon atom only (CO 1.17, NbC 2.53, MoC 2.02 A/ac>?;) and the other σ-bonded to Mo (MoC 1.944 A/ac>?;) and π-bonded to Nb (CO 1.22, NbC 2.22, NbO 2.26 A/ac>?;). Three types of carbonyl groups present in IV give rise to strong IR bands at 1870, 1700 and 1560 cm?1 assigned to the terminal, μ-bridging and σ, π-bridging CO groups respectively. Complex IV has a similar structure. The electronic structure of IV and its dissociation across the NbMo bond are discussed.  相似文献   

19.
In electrothermal atomic absorption spectrometry, the signals corresponding to total absorption (TA), background(BG) adn the required atomic absorption signal (TA - BG) show good temporal reproducibility. For improvement of the detection power, the time-resolved signals can be summed with the aid of a microcomputer. The summed signal height is exponentially related to the number (N) of measurements up to N = 40. The detection power is improved according to the Nsol12 law with a practical limit at summation of ca. 20 signals. Concentrations which are around the detection limit in single measurements (lead and cadmium in drinking water or lead in urine) can readily be determined by summation of signals from, for example, sixteen 20-μl injections of urine. Resulting high summed backgrounds, giving absorbances > 5, can be compensated.  相似文献   

20.
The average vicinal coupling constants of 1-bromo-2-chloroethane at different temperatures have been obtained in a gas-phase 1H NMR study of 1,2-disubstituted ethanes. Analysis of the experimental data, assuming a ”static” model with constant values for the vicinal coupling constants of the individual rotamers, yielded unacceptable results. A ”dynamic” model, which takes into account the torsional vibrations, has therefore been developed and used to analyse the gas-phase results. The values at 305 K for the vicinal coupling constants of the individual rotamers are: trans rotamer JT = 13.6 Hz, J't = 4.9 Hz; gauche rotamer 12(JG1 + JG2) = 0.9 Hz, 12(J'G1 + J'G2) = 7.8 Hz. These couplings are temperature-dependent with JT changing about four times as much as the other coupling constants. Analysis of the average vicinal coupling constants, measured in different solvents, with the “dynamic” model gave values for the vicinal coupling constants of the individual rotamers in excellent agreement with those from the gas-phase data. The value of 12(JG1+ JG2) decreases noticeably with increasing dielectric constant of the solution. This can be explained by changes in the average dihedral angle between the coupling protons, in the gauche rotamers, which leads to changes of JG1 and JG2 in the same direction.  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