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 464 毫秒
1.
Potential energy curves for the ground and some low energy excited states of a number of complexes with a 3d 5 electronic configuration have been computed from INDO type SCF MO calculations. The results agree extremely well with the known ground states of the complex ions MnF 6 4? , FeF 6 3? , CoF 6 2? , and Fe(CN) 6 3? , in particular the crossover from high to low spin being obtained for changes in both central metal ion oxidation state and ligand. The calculated contraction in metal ligand distance on passing from the high spin to the low spin state is ~ 0.05 Å for each complex in very good agreement with the value indicated by pressure dependent magnetic measurements. Computed electronic transition energies involving bothd-d type and charge-transfer excitations compare favourably with observed spectroscopic values.  相似文献   

2.
Metastable decay of (N2) n + , formed in a supersonic jet and ionized by electron impact, has been analyzed forn≤50. The probability for decay of (N2) n + into (N2) n?x + , plotted versusx, exhibits pronounced oscillations. The “period” of these oscillations increases with increasing precursor sizen, but converges to an average value of approximately 4.7 beyondn=25.  相似文献   

3.
Quantum chemical ab initio calculations have been performed for the vertical excitation energies and oscillator strengths of all low-lying electronically excited states of small helium cluster ions, He n + ,n=2, ..., 7. The geometrical structures of the ions were fixed at the equilibrium geometries of the respective ground states, for He 4 + and He 5 + also one alternative structure was considered. The low-lying excited states can be classified into two categories: the electronic transition can occur either within the central He 2 + or He 3 + unit or from the peripheral weakly bound He atoms to this unit. The latter transitions are very weak (f≈0.001), closely spaced, with vertical excitation energies of about 5.7 eV. The He 2 + and He 3 + units have strong transitions at 9.93 and 5.55 eV, respectively; these transitions are only slightly blue-shifted if He 2 + or He 3 + are placed as “chromophores” into the centre of a larger He n + cluster. The large difference in the vertical excitation energy of the strong transition should enable an experimental decision of the question whether the cluster ions have He 2 + or He 3 + cores.  相似文献   

4.
The vibrational structure of the first band of the photoelectron (PE) spectrum of HO 2 ? and DO 2 ? has been calculated on the basis of (slightly modified) ab initio potentials. The best agreement with the experimental spectrum of HO 2 ? is obtained for a vibrational temperature of ca. 600 K. “Peak D”, which has been under debate in earlier work, is composed of two transitions, with the “hot” transition 3 1 1 being more intense than the adiabatic transition. Since thev 2 bending mode of DO2 has significant OO stretching character, the vibrational structure of the PE spectrum of DO 2 ? is more complex than that of HO 2 ? . Large-scale RCCSD(T) calculations of the equilibrium electron affinity of HO2 yield 1.058 eV which agrees with the experimental value of 1.044 ± 0.020 eV.  相似文献   

5.
The difference in charge exchange rate in collisions between spin oriented sodium atoms and H 2 + ions has been measured at an energy of about 1 eV. H 2 + was stored in a Penning trap and polarized by spin exchange with Na beam atoms from a hexapole magnet. The ion loss from the trap due to charge exchange was different as we depolarized the atomic beam. From the data we obtain a ratio of cross sections for singlet and triplet collisionsQ 1/Q 3=1.5±0.2 andQ 3=1.2·10?15 cm2.  相似文献   

6.
Eleven isomers with the PyC2H 5 composition, which include three conventional (1–3) and eight distonic radical cations (4–11), have been generated and in most cases successfully characterized in the gas phase via tandem-in-space multiple-stage pentaquadrupole MS2 and MS3 experiments. The three conventional radical cations, that is, the ionized ethylpyridines C2H5-C5H4N (1–3), were generated via direct 70-eV electron ionization of the neutrals, whereas sequences of chemical ionization and collision-induced dissociation (CID) or mass-selected ion-molecule reactions were used to generate the distonic ions H2C·?C5H4N+?CH3 (4–6), CH3?C5H4N+?CH 2 · (7–9), C5H5N+?CH2CH 2 · (10), and C5H5N+?CH·?CH3 (11). Unique features of the low-energy (15-eV) CID and ion-molecule reaction chemistry with the diradical oxygen molecule of the isomers were used for their structural characterization. All the ion-molecule reaction products of a mass-selected ion, each associated with its corresponding CID fragments, were collected in a single three-dimensional mass spectrum. Ab initio calculations at the ROMP2/6–31G(d, p)//6–31G(d, p)+ZPE level of theory were performed to estimate the energetics involved in interconversions within the PyC2H5 system, which provided theoretical support for facile 4?7 interconversion evidenced in both CID and ion-molecule reaction experiments. The ab initio spin densities for the a-distonic ions 4–9 and 11 were found to be largely on the methylene or methyne formal radical sites, which thus ruled out substantial odd-spin derealization throughout the neighboring pyridine ring. However, only 8 and 9 (and 10) react extensively with oxygen by radical coupling, hence high spin densities on the radical site of the distonic ions do not necessarily lead to radical coupling reaction with oxygen. The very typical “spatially separated” ab initio charge and spin densities of 4–11 were used to classify them as distonic ions, whereas 1–3 show, as expected, “localized” electronic structures characteristic of conventional radical ions.  相似文献   

7.
The oscillator strengths of the “d-d” transitions of the ion [PtCl4]2? were calculated. The method, based on the evaluation of the MO of the distorted ion, gives results in good accordance with the experimental data. The transition to 1 B 1g (a 1g b 1g * ), as a consequence of the participation of the orbital 6s, mixed with \(5d_{z^2 } \) in \(a_{1_g } ,a_{1_g } \) is characterized by a very low intensity in the case of thex,y polarization.  相似文献   

8.
The dissociation mechanisms of dichlorocarbene dications following collisional activation have been investigated via tandem mass spectrometric techniques and semi-empirical calculations. Three channels appear to be significant: {fx1019-1} The second channel becomes dominant at high internal energy. Production of ground state fragments (channel 1) involves a transition driven by spin—orbit coupling from the CCl 2 2+ $CCl_2^2 \tilde X^1 \Sigma _g^ + $ state to the CCl 2 2? ā3Σ u ? state en route to the fragments. The dissociation barrier for the production of ground state fragments from the ground electronic state of CCl 2 2+ via the spin—orbit-induced transition is equal to 420 kJ mol?1. The dissociation pathway that corresponds to channel 3 includes a first isomerization step from the linear Cl-C-Cl2+ structure to a bent Cl-Cl-C2+ connectivity. The calculated isomerization barrier amounts to 550 kJ mol?1. The calculated reverse activation barriers are compatible with the measured kinetic energy released on the fragments.  相似文献   

9.
Guided ion beam mass spectrometry is used to measure the cross sections as a function of kinetic energy for reaction of SiH4 with O+(4S), O 2 + (2Πg,v=0), N+(3P), and N 2 + (2Σ g + ,v=0). All four ions react with silane by dissociative charge-transfer to form SiH m + (m=0?3), and all but N 2 + also form SiXH m + products where (m=0?3) andX=O, O2 or N. The overall reactivity of the O+, O 2 + , and N+ systems show little dependence on kinetic energy, but for the case of N 2 + , the reaction probability and product distribution relies heavily on the kinetic energy of the system. The present results are compared with those previously reported for reactions of the rare gas ions with silane [13] and are discussed in terms of vertical ionization from the 1t 2 and 3a 1 bands of SiH4. Thermal reaction rates are also provided and dicussed.  相似文献   

10.
11.
The absolute cross section for photodissociation of Ar2N 2 + was measured as a function of wavelength in the 470–550 nm range. A structureless broad band was observed; the cross section has a maximum of ~ 210 × 10?18 cm2 at ~ 500 nm. The measurement of the photofragment time-of-flight spectrum shows that(1) N 2 + , Ar+ and Ar 2 + are produced in the photodissociation of Ar2N 2 + in the wavelength range studied, and that(2) the observed visible absorption band is ascribable to a parallel-type transition of Ar2N 2 + , which possibly retains a linear geometry.  相似文献   

12.
13.
Optical absorption spectra of cobalt cluster ions, Co n + , and vanadium cluster ions, V n + , were analyzed by a theoretical calculation based on the spin-polarized DV- method, and their electronic and geometric structures were obtained. Relative absorption cross section associated with each electronic transition was calculated; the calculation enables a qualitative comparison of calculated spectrum with a measured one not only in its transition energy but also in its intensity profile. This analysis shows that Co 4 + , Co 3 + , and V 4 + have, respectively, a tetrahedral structure with a bond distance of 2.00Å, an equilateral triangle with a bond distance of 2.30Å, and a distorted tetrahedral structure with five bonds having a distance of 2.34 Å and one of 2.89Å. The differences in the population between majority and minority spins (spin-difference) evaluated from the electronic structure thus obtained were 2.0, 1.7, and zero per atom in Co 3 + , Co 4 + , and V 4 + , respectively. These spin differences indicate a ferromagnetic and an antiferromagnetic spin-coupling in the cobalt and vanadium cluster ions, respectively.  相似文献   

14.
By treatment of 1,3-bis(3,4-dimethoxybenzyl)-3,4,5,6-tetrahydropyrimidinium chloride (1) with KOBu t and [PtCl2(PEt3)2]2 N-coordinated platinum complex (2) is obtained. The Pt atom is coordinated in square planar arrangements by two chloride ions in a trans-configuration, the N-formyl-N,N′-bisaryltrimethylenediamine nitrogen atom, and the phosphine P atom. An extensive three-dimensional network of three C-H…O hydrogen bonds, two C-H…π and one π…π interactions are responsible for the crystal stabilization. Intermolecular hydrogen bonds and C-H…π interactions produce R 2 2 (6), R 2 2 (22), R 2 2 (24), R 3 3 (23), R 4 4 (26), and R 4 4 (32) rings.  相似文献   

15.
Photofragment spectroscopy of N 2 + has been studied in the wavelength range 343–404 nm using an excimer-pumped dye laser with a spectral resolution of 0.2 cm?1. The observed bands are assigned to transitions from thev″=23?26 levels of theX 2Σ g + state to highlying rovibrational levels (v′≈46–48) of theB 2Σ u + state, forming quasibound (predissociating) states above the dissociation limit N+(3 P)+N(4 S 0). Measurement of the photofragment kinetic energies allows to establish an absolute energy scale for the transitions with respect to the dissociation limit. Molecular constants for the lower and upper states of the observed transitions are determined. The measurements allow the first direct determination of the N 2 + dissociation energyD 0 0 (N 2 + ). Some high-resolution (0.04 cm?1) measurements show the fine-structure splitting and lifetime broadening of the excitation lines.  相似文献   

16.
The oxygen nonstoichiometry δ of lanthanum cobaltite doped with acceptor impurities (Sr and Ni), La1 ? x SrxCo0.9Ni0.1O3 ? δ (x = 0.1, 0.3), was studied by high-temperature thermogravimetry over the temperature and pressure ranges 723 K ≤ T ≤ 1373 K and 10?3 atm ≤ $p_{O_2 } $ ≤ 1 atm. The partial replacement of cobalt with nickel and lanthanum with strontium increased the oxygen nonstoichiometry δ. The partial molar enthalpies $\Delta \bar H^\circ _O $ and entropies $\Delta \bar S^\circ _O $ of solution of oxygen in the solid phase were calculated. Models of point defect formation were suggested and analyzed. The equilibrium constants of formation and concentrations of predominant point defects, ionized oxygen vacancies V o .. , holes Me Co . (Co Co . and Ni Co . ), and electrons Me Co (Co Co and Ni Co ) localized on 3d transition metals, were determined by nonlinear regression from the experimental and theoretical logp $p_{O_2 } $ ?δ dependences.  相似文献   

17.
Calorimetric methods were used to study the enthalpy of solution of the alkaloids glaucine C21H25NO4 and lappaconitine C32H44N2O8 in 96% ethanol and their heat capacity in the temperature ranges 173–298.15 and 173–423 K. Equations describing the dependences ΔH sol m f( $\sqrt m $ ) (m is molal concentration) and C p 0 f( $\sqrt T $ ) for the compounds under study were derived. The enthalpies of combustion, melting, and formation of C21H25NO4 and C32H44N2O8 were estimated by means of theoretical calculations.  相似文献   

18.
Various factors influencing the performance of a Hadamard transform time-of-flight mass spectrometer (HT-TOFMS) have been investigated. Using a nitrogen corona discharge to produce an ion stream of N 2 + , N 3 + , and N 4 + , it is found for spectra containing only N 4 + that the signal-to-noise ratio (SNR) closely approaches the value calculated from the ion background by assuming that the ion background follows a Poisson distribution. In contrast, for a more intense beam containing N 2 + , N 3 + , and N 4 + , the SNR is less than its theoretical value because of the appearance of discrete spikes in the mass spectrum caused by deviations in the actual modulation sequence from the ideal one. These spikes can be reduced, however, by decreasing the modulation voltage. Under these optimized conditions, the pseudo-random sequence length is varied to understand how it alters SNR, mass resolution, and scan speed. When the length of the pseudo-random sequence is doubled, the SNR increases by √2 while the time necessary to record a mass spectrum also doubles. Mass resolution can be varied between 500 and 1200 at m/z = 609 as the sequence length, modulation speed (10 MHz, 25 MHz), and acquisition rate (up to 50 MHz) are changed. Scan speeds of 6000 passes per s can be obtained using a sequence containing 4095 elements modulated at 25 MHz. The capability to tailor the HT-TOFMS to increase the scan speed and resolution with a constant 50% duty cycle makes the technique extremely appealing as a mass analyzer for measuring rapid changes in the composition of an ion stream.  相似文献   

19.
Reactivity of positively charged cobalt cluster ions (Co n + ,n=2?22), produce by laser vaporization, with various gas samples (CH4, N2, H2, C2H4, and C2H2) were systematically investigated by using a fast-flow reactor. The reactivity of Co n + with the various gas samples is qualitatively consistent with the adsorption rate of the gas to cobalt metal surfaces. Co n + highly reacts with C2H2 as characterized by the adsorption rate to metal surfaces, and it indicates no size dependence. In contrast, the reactions of Co n + with the other gas samples indicate a similar cluster size dependence; atn=4, 5, and 10?15, Co n + highly reacts. The difference can be explained by the amount of the activation energy for chemisorption reaction. Compared with neutral cobalt clusters, the size dependence is almost similar except for Co 4 + and Co 5 + . The reactivity enhancement of Co 4 + and Co 5 + indicates that the cobalt cluster ions are presumed to have an active site for chemisorption atn=4 and 5, induced by the influence of positive charge.  相似文献   

20.
Semiempirical self-consistent-field molecular-orbital calculations are carried out for six iron-pentacyanide complexes and are used to interpret their experimental Mössbauer quadrupole splittings. Probable orientations are identified for the C6H 5 ? and NO 2 ? groups in Fe(CN)5NOC6H 5 ?3 and Fe(CN)5NO 2 ?4 . Calculations on Fe(CN)5NO?2 and Fe(CN)5NO?3 can simultaneously be brought into agreement with experiment by reparametrization to make the NO group more positively charged. All the calculations indicate the importance of including all the Fe 3d and 4p orbitals in the calculations and of considering neighboring-atom effects.  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