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.
A series of cluster anions C n X? are produced from laser ablation of appropriate samples, where X is selected as a group-VB element. The recorded mass spectra of these cluster anions display a drastic even/odd alternation on ion intensities: For C n N?, only anions with oddn can be observed; For C n P? and C n As?, cluster anions with evenn are produced but with lower signal intensities; For C n Sb?, the signal intensity of clusters does not show even/odd alternation; Finally, for C n Bi?, the intensities of cluster anions with evenn are higher than those with oddn. This parity effect can be attributed to the linear structure of the cluster anions, and the parity reversal of C n X? from C n N? to C n Bi? can be explained from the electronegativity decreasing of the heteroatom X as it descends in the group. The Hückel model was applied to account the structural feature of these clusters.  相似文献   

2.
In an attempt to improve the understanding of the electronic structure of the Ni2+ thiocarboxylates, we have analized, as a zeroth-order approximation, the electronic structure of the Ni4+2 dimer in vacuo. Two small-size Slater-type orbitals (STO) basis sets have been used in order to study basis effects, referring particularly to the details of the d-d interactions. All the electrons have been included and all the molecular integrals have been computed accurately. All the multiplets of the {Ar2}σ2gσ2gπ4gπ4uδxgδ4?xu (x = 0 to 4) configurations have been calculated at different values of the Ni2+Ni2+ distance, R, ranging from 3.20 to 5.20 a.u. in the smaller basis calculation and from 3.20 to 4.20 a.u. in the larger basis calculation. Inclusion of configuration interaction (CI) limited to the δxgσ4?xu (x = 0 to 4) configurations yields a 3Σ?g as the lowest multiplet. This CI reduces Re from 2.03 to 1.76 Å, an effect also encountered in much more extensive CI calculations on Ni2. The bonding in Ni4+2 is discussed in terms of the R dependence of the orbital energies; σ, π, and δ interactions seem to make important contributions to the bonding, in view of their orbital splittings and stabilizations with respect to the free-ion values. The present results are compared with previous nonempirical studies on the Ni2 and Ni+2, as well as with empirical relationships among different molecular constants. The relevance of the results on Ni4+2 to the question of the metal-metal interaction in {Ni(RCOS)2}2 compounds is difficult to discuss with certainty, but, mainly from the theoretical geometry obtained, it is argued that the Ni2+Ni2+ interactions might have a very small contribution to the selling of the electronic structure of the thiocarboxylates.  相似文献   

3.
a digital simulation analysis is presented of the deleterious effects of uncompensated solution resistance, Rus, on the evaluation of standard rate constant, ksob, by cyclic voltammetry. The results are expressed in terms of systematic deviations of “apparent measured” rate constants, ksob(app), evaluated in the conventional manner without regard for Rus, from the corresponding actual values, ksob(true), as a function of Rus and other experimental parameters. Attention is focused on the effects of altering the electrode area and the double-layer capacitance on the extent of the deviations between ksob(app) and ksob(true), and on comparisons with corresponding simulated results obtained from phase-selective a.c. impedance data. The extent to which ksob(app) <ksob(true) for small Rus values was found to be similar for the cyclic and a.c. voltammetric techniques. The latter method is, however, regarded as being preferable under most circumstances in view of the greater ease of minimising, as well as evaluating, Rus for a.c. impedance measurements. The influence of solution resistance on ksob measurements with microelectrodes and without iR compensation is also considered.  相似文献   

4.
5.
A new mechanism Of H2 dissociation in electrical discharges (1011 ? ne ? 1012 cm?3, 2.10?16 ? E/N ? 3.10?16 V cm2, 300 ? Tg ? 1000 K, 3 ? p ? 30 torr) is presented and discussed. In this mechanism, called joint vibro-electronic mechanism (JVE), the electrons of the discharge create a strong vibrational disequilibrium with respect to the gas temperature (Tg) and promote electronic transitions from all vibrational levels of 1Σg H2 state to the repulsive 3Σu one. Moreover the V-V (vibration-vibration) and V-T (vibration-translation) energy exchanges are considered for building up the vibrational distribution of 1Σg state. A complete set of e - D cross sections (e + H2(1Σg,ν) → e + H2 (3Σu) → + 2H, ν = 0,14) is calculated by using an extension of the semiclassical Gryzinski theory in combination with the Franck-Condon principle. Dissociation rates calculated according to JVE are larger either than those obtained by the pure vibrational mechanism (PVM) discussed in our previous work or than those from the direct electronic impact mechanism (DEM) from the ground vibrational level. The behaviour of JVE rates as a function of gas temperature (Tg), of E/N, of electron density (ne) and of pressure is then reported. The results show strong differences as compared, with the corresponding values obtained, with PVM. Finally the influence of the atoms as well as their recombination on the dissociation rates is discussed. The results have been obtained by solving a system of vibrational master equations.  相似文献   

6.
Gaseous mixtures of phosphine and germane have been investigated by ion trap mass spectrometry. Reaction pathways together with rate constants of the main reactions are reported. The mechanisms of ion/molecule reactions have been elucidated by single and multiple isolation steps. The GeHn+ (n = 1–3) ions react with phosphine to give GePHn+ (n = 2–4) ions. The GePH4+ ion further reacts with GeH4 to yield Ge2PH6+. The GePHn+ (n = 2–4) mixed ionic family also originates from the P+ phosphine primary ion, as well as from the P2Hn+ (n = 0–3) secondary ions of phosphine reacting with neutral germane and from Ge2H2+ reacting with phosphine. The main reaction pathways of the PHn+ (n = 0–2) ions with GeH4 lead to the formation of the GeH2+ and GeH3+ ionic species. Protonation of phosphine from different ionic precursors is a very common process and yields the stable phosphonium ion, PH4+. Trends in total abundances of secondary GePHn+ (n = 2–4) ions as function of reaction time for different PH3/GeH4 pressure ratios show that excess of germane slightly affects the nucleation of mixed Ge-P ions.  相似文献   

7.
The structure of barium-titanium-metaborate xBaO-xB2O3-yTiO2 (y=0%, 4%, 8%, 16% and x=50-y/2) amorphous and crystallized powders, obtained using a polymeric precursor method, was investigated by Ti and B K-edge X-ray absorption spectroscopy (XAS) and 11B-NMR high-resolution techniques. XANES study of amorphous samples shows that Ti4+ ions exist as [4]Ti species associated to [6]Ti and [5]Ti species in a practically equivalent amount. After crystallization, titanium environment is predominately composed by [6]Ti species. According to XANES results obtained at the B K-edge, the fraction of boron in tetrahedral sites ([4]B) reduces as the amount of TiO2 is increased from x=0% to 4%, with a consequent increase of boron in trigonal sites ([3]B). By a combination of 11B-NMR spin-echo and triple quantum magic angle spinning (3Q-MAS) techniques, the detailed borate speciation was determined as consisting in [4]B and two kind of trigonal sites, [3]BA and [3]BB, corresponding, respectively, to borates sharing three and two O atoms with other boron units. NMR results reveal not only the reduction in boron coordination also seen by XANES but also the simultaneous reduction in the condensation degree of trigonal units, when the Ti content is increased in the glass. In crystallized samples, β-BaB2O4 and BaTi(BO3)2 phases were identified and quantified by 11B-NMR.  相似文献   

8.
The range of chemical flexibilities of the hexagonal frameworks (Ta6Si4O26)6? and (Ta14Si4O47)8? have been partially explored. This has been done with high-temperature preparations as in general ionic mobilities in these frameworks are too low to permit low-temperature ion exchange. Ionic site potential calculations indicate that preferential site-occupancy factors as well as geometric constraints are responsible for the absence of ionic motion. New phases K6?xNaxTa6Si4O26 (x ? 4), K8?xNaxTa14Si4O47 (x ? 5), and impure Ba3?xNa2xTa6Si4O26 have been prepared. Introduction of up to 2 moles of Li+ and 1 mole of Mg2+ ions per formula unit into sites of the framework not normally occupied has been demonstrated as well as the possibility of partially substituting Zr4+ for Ta5+ ions. Substitutions designed to introduce large tunnel vacancies in the presence of only monovalent K+ or Na+ ions (P for Si, W for Ta and F for O) generally proved unsuccessful. Competitive phases also frustrated attempts to substitute either the larger Rb+ or the smaller Li+ ions into the large-tunnel sites. A large area of solid solution was discovered in the BaONa2OTa2O5 phase diagram; it has a (TaO3)-framework with the structure of tetragonal potassium tungsten bronze.  相似文献   

9.
Ab initio calculations at SCF and CEPA levels using large Gaussian basis sets have been performed for the two lowest electronic states,X 2 Σ+ andA 2 Π, of HeAr+. Spin-orbit coupling (SOC) effects have been added using a semiempirical treatment. The resulting potential curves for the three statesX,A 1, andA 2 have been used to evaluate molecular constants such as vibrational intervals ΔG(v + 1/2) and rotational constantsB v as well as — by means of a Dunham expansion — equilibrium constants such asR e , ω e ,B e etc. Comparison with the experimental data from UV emission spectroscopy shows that the calculated potential curves are slightly too shallow and have too large equilibrium distances:D e = 242 cm?1 andR e = 2.66 Å compared to the experimental values of 262 cm?1 and 2.585 Å, respectively, for theX 2Σ+ ground state. However, the ab initio calculations yield more bound vibrational levels than observed experimentally and allow for a more complete Dunham analysis, in particular for theA 2 state. The experimental value of 154 cm?1 for the dissociation energyD e of this state is certainly too low; our best estimate is 180±5 cm?1. For theA 1 state our calculations are predictions since this state has not yet been observed experimentally.  相似文献   

10.
A simple experimental arrangement was applied for the measurement and the evaluation of pitting corrosion currents operating under natural conditions. The feasibility of the procedure was examined by using Zn as a test metal, K2CrO4, Na2HPO4 and Na2WO4 as inhibitors, and Cl?, Br? and I? as pitting corrosion agents. Both the type and concentration of the inhibiting and corroding agents were varied in a programmed manner. In CrO42? and HPO42? solutions, the pitting corrosion currents started to flow after an induction period, which decreased with increase in the concentration of the attacking agent. In WO42? solutions, on the other hand, initially high currents were recorded due to the reduction of the agent to soluble, non-inhibiting species.In all solutions tested the corrosion current reached steady-state values which depended on the type and concentration of both the inhibiting and the aggressive anions. When keeping the inhibitor concentration constant, the corrosion current varied with the concentration of the aggressor according to: log icorr = a1 + b1 log cagg On the other hand, in solutions of constant aggressor concentration, with varying inhibitor amounts, the relation was: log icorr = a2 ? b2 log cinh where a1(a2) and b1(b2) are constants.The two equations were derived theoretically on the basis of competitive adsorption of the two counteracting agents on the surface of the metal. Comparison between the experimental values of a and b, with the corresponding terms of the theoretical equations was made.The aggressivity of the three tested anions decreases in the order Cl? > Br? > I?, whilst inhibition varied as CrO42? > HPO42? > WO42?.  相似文献   

11.
The ribbed functionalization of the clathrochelate iron(II) tris-dioximates as a potential “molecular scaffold” for the synthesis of polyfunctional and polytopic complexes with closo-dodecaborate-anion substituents was performed. closo-Dodecaborate-substituted clathrochelate [FeBd2(Cl(B12H11NH)Gm)(BF)2]2? (where Bd2? is α-benzyldioxime dianion, Gm is glyoxime residue) dianion was prepared starting from dichloride clathrochelate FeBd2(Cl2Gm)(BF)2 precursor with amino-closo-dodecaborate (NBu4)[B12H11NH3] in the presence of potassium tert-amylate. This clathrochelate dianion was isolated as a tetra-n-butylammonium salt and characterized using elemental analysis, MALDI-TOF and PD mass, IR, UV-Vis, 57Fe Mössbauer spectra as well as by 1H, 11B and 13C{1H} NMR spectra.  相似文献   

12.
“Pd0(PPh3)2” has been generated by electroreduction of Cl2Pd(PPh3)2 in the absence of PPh3. Its structure is consistent with [ClxPd0(PPh3)2]nnx, with values of x and/or n depending on the chloride ion concentration. The half-lives of its reactions with PhI or PhBr have been determined, and found to be 350 and 100 times, respectively, as large as those for Pd(PPh3)4.  相似文献   

13.
The reactions of a range of 2-arsa- and 2-stiba-1,3-dionato lithium complexes with group 4-7 metals have been investigated. These have given rise to several complexes in which an arsadionate acts as a chelating ligand; [V{η2-O,O-OC(But)AsC(But)O}3], [M{η2-O,O-OC(But)AsC(But)O}2(DME)], M=Cr or Mn; or as an η1-As-diacylarsenide, [MnBr(CO)4{As[C(O)But]2Li(DME)}]2. In addition, reactions of lithium arsadionates with TaCl5 have led to metal mediated arsadionate decomposition reactions and arsadionate oxidative coupling reactions to give the known arsaalkyne tetramer, As4C4But4, and the new tetraacyldiarsane, [{As[C(O)Mes]2}2] Mes=mesityl, respectively. The treatment of several lithium arsadionates with [MoBr2(CO)2(PPh3)2] has also initiated arsadionate decomposition reactions and the formation of the metal carboxylate complexes, [MoBr(CO)22-O2C(R)}(PPh3)2] R=But, Ph, Mes. The X-ray crystal structures of six of the prepared complexes are discussed.  相似文献   

14.
CpZrCl3·dme was treated with Na[El(OtBu)3], El = Ge, Sn, Pb, respectively. The addition of Na[Sn(OtBu)3] to CpZrCl3·dme caused rapid cyclopentadienide loss and the equally rapid appearance of CpSnCl, half of which crystallized as the trinuclear complex {[ZrCl(OtBu)3]2·CpSnCl}. Pristine CpSnCl reacted almost instantly with NaOtBu to give NaCp and Na[Sn(OtBu)3], which co-crystallized as a coordination polymer. Na[Ge(OtBu)3] also displaced Cp from zirconium, but with a different product distribution, giving Cp2Ge, fac-[Ge(μ-tBuO)3ZrCl(OtBu)2], and ZrCl(OtBu)3. By contrast, Na[Pb(OtBu)3] only exchanged its tert-butoxide groups with zirconium to furnish CpZr(OtBu)3 and PbCl2. The solid-state structures of {[ZrCl(OtBu)3]2·CpSnCl}, fac-[Ge(μ-tBuO)3ZrCl(OtBu)2], and {NaCp·Na[Sn(OtBu)3]}n were determined.  相似文献   

15.
The OPAL research reactor in Australia has been used to determine k 0 values for 134mCs, 134Cs, 192Ir and 194Ir. Values for 24Na have also been measured for quality control. The neutron flux at the irradiation positions was very highly thermalised (f > 2,000), resulting in almost negligible activation by epithermal neutrons. As a consequence, the contribution to the total uncertainty of the k 0 values from epithermal-related factors such as Q 0 and $ \bar{E}_{\text{r}} $ was very small. The measured caesium k 0 values have been compared with the library values as well as with recent measurements by St Pierre et al. and Farina Arboccò et al. While there are k 0 values for 194Ir in the library, no 192Ir values have been measured previously. Despite 192Ir having a higher sensitivity than 194Ir, k 0 values were not measured during the establishment of the k 0-method because the nuclear data available at the time indicated that the activation cross-section of 191Ir deviated significantly from 1/v behaviour (g(T n ) ≠ 1), which would result in unacceptable errors if k 0 analysis were to be carried out using the Høgdahl convention. However later nuclear data compilations showed that 191Ir has better 1/v behaviour than previously reported, making it suitable for k 0 analysis using the Høgdahl convention. For completeness, k 0 values have been determined using both the Høgdahl and modified-Westcott conventions and these have been compared with library (194Ir) and calculated values.  相似文献   

16.
The solid state synthesis of Cs4Nb6Fi8.5Ii3.5Ia6 starting from Nb6F15 binary fluoride, as well as its crystal structure determined by X-ray single crystal diffraction, are presented in this work. This novel cluster compound is based on a Nb6Ii3Fi6Li3Ia6 (L=F, I) discrete unit and crystallizes in the monoclinic system (space group C2/m; Z=4 ; a=10.4363(4) Å, b=18.1227(7) Å, c=19.5102(9) Å β=101.223(1)°, V=3619.5(3) Å3, R1=0.057; wR2=0.159). This halide is the first octahedral niobium cluster compound containing unshared terminal Ia ligands together with ordered μ2-Ii and μ2-Fi ligands on nine inner positions whilst the three last ones (Li) are slightly affected by a I/F random occupancy. The structural findings are discussed and compared with those of Nb6F15, Nb6I11, CsNb6I11 and the fluorochlorides and fluorobromides recently reported.  相似文献   

17.
The crystal structures of M+VO3(M+ = K, NH4, and Cs) have been refined using three-dimensional counter-diffractometer X-ray data and full-matrix least-squares methods. The structure of these compounds is characterized by a (V5+O2?3)? chain extending along the c-axis (Pbcm orientation), with adjacent chains linked by the alkali metal cation. The structure may be considered as a variant of the pyroxene structure, and standard atom nomenclature is proposed in order to facilitate comparison with silicate pyroxenes. Structural variation across this series is discussed in detail and is compared with the analogous M+M3+Si2O6 (M+ = Li, Na; M3+ = Al, Cr, Fe, Sc, In) series.  相似文献   

18.
The behaviour in aqueous solution of some aminocarboxylates of the type NH2(CH2)nS(CH2)m-1COO? (abbreviated as n,m-NSO; n and m = 2 or 3) in equilibria with H+, Cu2+ and Ni2+ ions has been investigated potentiometrically and calorimetrically at 25° C in a 0.5 M KNO3 medium.The protonation of the amino function and especially the carboxylate function is attended by strong desolvation effects, which are characterized by low exothermic enthalpies and strongly positive entropies of protonation.In [Cu(n,m-NSO)]+ and [Ni(n,m-NSO)]+ the aminocarboxylates act as tridentate ligands, forming complexes with a strong hard—hard character. The biligand species [Ni(n,m-NSO)2] behave as six-coordinated complexes whereas in [Cu(n,m-NSO)2] the second ligand is bound only through the N and S donor, forming a five-coordinated species.Finally, the n,m-NSO ligands also form protonated species with the Cu2+ ion.  相似文献   

19.
Emulsion membrane systems consisting of (1) an aqueous source phase containing 0.001 M Cd (NO3)2 and/or 0.001 M AgNO3 and varying concentrations of SCN, (2) a toluene membrane containing dicyclohexano-18-crown-6 (DC18C6) (0.02 M) and the surfactant Span 80 (sorbitan monooleate) (3% v/v), and (3) an aqueous receiving phase containing MgS2O3 or Mg (NO3)2 were studied with respect to the disappearance of Cd (II) and/or Ag (I) from the source phase as a function of time. The transport rate of Cd (II) was highest when a maximum amount of the Cd(II) in the source phase was present as Cd(SCN)2 ([SCN] =0.4 M). Cadmium(II) was transported over Ag(I), which is present mainly as Ag(SCN)43−, by 5- and 55-fold in 5 minutes with 0.4 M SCN in the source phase and 0.3 M S2O2−3 and 0.3 M NO3, respectively, in the receiving phase. In these competitive experiments, the total percent of Cd (II) transported was 98 and 84, respectively. The results are explained using the various equilibrium constants for cation-DC18C6, cation-SCN and cation-S2O2−3 interactions. These results indicate that rapid transport occurs when a cation is present in the source phase as a neutral complex. Selectivity for neutral species can be designed into these membrane systems when other cations interact with the source-phase anion to form charged species. Emulsion systems like those above were studied with respect to the appearance of Li+ in the source phase as a measure of membrane breakage. Maximum membrane stability was obtained when the ionic strengths of the source and receiving phases were equal.  相似文献   

20.
Investigation of π-(CH3nC5H5–nMn(CO)3 (Men-CpMnT) n = 0−5 and t-butyl CpMnT mass spectra showed that MenCpMnT molecular ion (M+) fragmentation occurs by a simpler scheme than that for MenCpReT M+ molecular ions. The reason is that MnCp and MnCO bonds are not as strong as the ReCp and ReCO bonds, and the relative “inertness” (compared to Re) of the Mn atom (ion), coordinated to the methylcyclopentadienyl ligand. Variations of M+ molecular ion intensity with different values of n are probably due to a complexity of electronic and spatial methyl-carbonyl group interactions in M+.  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