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 46 毫秒
1.
The Raman and infrared spectra of compounds NaLnTiO4 (Ln = lanthanide, including yttrium) are reported and discussed. Their most striking feature is a strong band in both spectra at about 900 cm?1. This band is ascribed to a vibration localized in the TiO bond directed towards the NaO layers. The relevant oxygen anion is very poorly charge compensated, and the TiO bond is, therefore, very strong. Pauling's electrostatic valence rule appears to be of great use in these considerations. These compounds do not show ferroelectricity.  相似文献   

2.
The electron density near the lithium nucleus in the species LiH, LiH+, Li2, Li2+, LiH2+, and Li2H+ was analyzed by transforming the SCF molecular orbitals into a sum of atomic contribnutions, for both core and valence orbitals. These “hybrid-atomic” orbitals were used to compare: electron densities, orbital polarizations, and orbital mean kinetic energies with the corresponding lithium atom quantities. Core-orbital electron densities at the lithium nucleus were observed to increase by up to 0.5% relative to the lithium atom 1s orbital. Lithium cores also exhibited polarization but, surprisingly, in the direction away from the internuclear region. Similar dramatic changes were seen in the electron densities of the valence orbitals of lithium: The electron density at the nucleus for these orbitals increased two-fold for homonuclear species and twenty-fold for heteronuclear triatomic species relative to the electron density at the nucleus in lithium atom. The polarization of the valence orbital electronic charge, in the vicinity of the lithium nucleus, was also away from the internuclear region. The mean “hybrid-atomic” orbital kinetic energies associated with the lithium atom in the molecules also showed changes relative to the free lithium atom. Such changes, accompanying bond formation, were relatively small for the lithium core orbitals (within 0.2% of the value for lithium atom). The orbital kinetic energies for the lithium valence electrons, however, increased considerably relative to the lithium atom: By a factor of about 2 in homonuclear diatomics, by a factor of 7 in heteronuclear diatomics, and by a factor of 11 in the triatomic species. In summary, the total electronic density (core plus valence) at the lithium nucleus remained remarkably constant for all of the species studied, regardless of the effective charge on lithium. Thus, the drastic changes noted in the individual lithium orbitals occurred in a cooperative fashion so as to preserve a constant total electron density in the vicinity of the lithium nucleus. In all cases, bond formation was accompanied by an increase in the orbital kinetic energy of the lithium valence orbital. We suggest that these two observations represent important and significant features of chemical bonding which have not previously been emphasized.  相似文献   

3.
Two anilinosalen and a mixed phenol‐anilinosalen ligands involving sterically hindered anilines moieties were synthesized. Their nickel(II) complexes 1 , 2 , and 3 were prepared and characterized. They could be readily one‐electron oxidized (E1/2=?0.30, ?0.26 and 0.10 V vs. Fc+/Fc, respectively) into anilinyl radicals species [ 1]+ , [ 2]+ , and [ 3]+ , respectively. The radical complexes are extremely stable and were isolated as single crystals. X‐ray crystallographic structures reveal that the changes in bond length resulting from oxidation do not exceed 0.02 Å within the ligand framework in the symmetrical [ 1]+ and [ 2]+ . No quinoid bond pattern was present. In contrast, larger structural rearrangements were evidenced for the unsymmetrical [ 3]+ , with shortening of one Cortho? Cmeta bond. Radical species [ 1]+ and [ 2]+ exhibit a strong absorption band at around 6000 cm?1 (class III mixed valence compounds). This band is significantly less intense than [ 3]+ , consistent with a rather localized anilinyl radical character, and thus a classification of this species as class II mixed‐valence compound. Magnetic and electronic properties, as well as structural parameters, have been computed by DFT methods.  相似文献   

4.
Composite photocatalyst films have been fabricated by depositing BiVO4 upon TiO2 via a sequential ionic layer adsorption reaction (SILAR) method. The photocatalytic materials were investigated by XRD, TEM, UV/Vis diffuse reflectance, inductively coupled plasma optical emission spectrometry (ICP‐OES), XPS, photoluminescence and Mott–Schottky analyses. SILAR processing was found to deposit monoclinic‐scheelite BiVO4 nanoparticles onto the surface, giving successive improvements in the films′ visible light harvesting. Electrochemical and valence band XPS studies revealed that the prepared heterojunctions have a type II band structure, with the BiVO4 conduction band and valence band lying cathodically shifted from those of TiO2. The photocatalytic activity of the films was measured by the decolourisation of the dye rhodamine 6G using λ>400 nm visible light. It was found that five SILAR cycles was optimal, with a pseudo‐first‐order rate constant of 0.004 min?1. As a reference material, the same SILAR modification has been made to an inactive wide‐band‐gap ZrO2 film, where the mismatch of conduction and valence band energies disallows charge separation. The photocatalytic activity of the BiVO4–ZrO2 system was found to be significantly reduced, highlighting the importance of charge separation across the interface. The mechanism of action of the photocatalysts has also been investigated, in particular the effect of self‐sensitisation by the model organic dye and the ability of the dye to inject electrons into the photocatalyst′s conduction band.  相似文献   

5.
SCF-Xα scattered wave cluster MO calculations for the oxyanions CuO?64 (D4h symmetry) and ZnO?64 (Td symmetry) yield results in good agreement with the X-ray photoelectron and X-ray emission spectra of CuO and ZnO, respectively. Agreement of the calculations with optical data is fair. Calculations of the valence electron and core electron hole states of these oxyanions support the assignment of photoelectron shakeup satellites to valence band to conduction band transitions. Calculated shakeup energies for the Cu2p core spectrum in CuO are 7.4 and 9.9 eV (cf. experimental values of 7.5 and 10.0 eV) while shakeup peaks in the valence region spectrum are predicted at 6.1 and 8.0 eV. (Cf. a broad peak with maximum at 8.1 eV observed experimentally.) The absence of intense low energy satellites in the spectra of ZnO is explained by the small amount of electron reorganization in the outer valence levels attendant upon hole formation.  相似文献   

6.
The electron density of tetragonal ZnP2 and Cdh2 has been calculated by the pseudopotential method. The valence band of ZnP2 and CdP2 consists of 48 subbands grouped into 4 subbands. The Zn(Cd)–P bond proved to be ionic-covalent, with a bond charge of 0.01(2)e whose maximum is shifted toward phosphorus. The (r) maximum of the shortest P1–P3 bond lies in its middle; the bond charge values are 0.25e (ZnP2) and 0.50e (CdP2). The charge maximum on the P2–P3 bond lies near the middle of the bond; it slightly deviates from the bond direction in the (P1, P2 P3) plane, indicating anisotropy of the bond. The partial electron densities have been calculated for the four subbands of the valence band. The reason for the splitting of the phosphorus s-band into two bands in ZnP2 and CdP2 was found to be sp-hybridization in the region of the s-band, which is the strongest for the shortest P1–P3 bond.Kemerovo State University. Translated fromZhurnal Strukturnoi Khimii, Vol. 34, No. 5, pp. 52–56, September–October, 1993.Translated by L. Smolina  相似文献   

7.
The ternary hafnium silicon arsenide, Hf(SixAs1−x)As, has been synthesized with a phase width of 0.5?x?0.7. Single-crystal X-ray diffraction studies on Hf(Si0.5As0.5)As showed that it adopts the ZrSiS-type structure (Pearson symbol tP6, space group P4/nmm, Z=2, a=3.6410(5) Å, c=8.155(1) Å). Physical property measurements indicated that it is metallic and Pauli paramagnetic. The electronic structure of Hf(Si0.5As0.5)As was investigated by examining plate-shaped crystals with laboratory-based X-ray photoelectron spectroscopy (XPS) and synchrotron radiation photoemission spectroscopy (PES). The Si 2p and As 3d XPS binding energies were consistent with assignments of anionic Si1− and As1-. However, the Hf charge could not be determined by analysis of the Hf 4f binding energy because of electron delocalization in the 5d band. To examine these charge assignments further, the valence band spectrum obtained by XPS and PES was interpreted with the aid of TB-LMTO band structure calculations. By collecting the PES spectra at different excitation energies to vary the photoionization cross-sections, the contributions from different elements to the valence band spectrum could be isolated. Fitting the XPS valence band spectrum to these elemental components resulted in charges that confirm that the formulation of the product is Hf2+[(Si0.5As0.5)As]2−.  相似文献   

8.
The electronic characteristics of mixed‐valence complexes are often inferred from the shape of the inter‐valence charge transfer (IVCT) band, which usually falls in the near infrared (NIR) region, and relationships derived from Marcus‐Hush theory. These analyses typically assume one single, dominant molecular conformation. The NIR spectra of the prototypical delocalised (Class III Robin–Day mixed‐valence) complexes [{Ru(pp)Cp’}2(μ‐C≡C?C≡C)]+ ([ 1 ]+: Cp’=Cp, pp=(PPh3)2; [ 2 ]+: Cp’=Cp, pp=dppe; [ 3 ]+: Cp’=Cp*, pp=dppe) feature a ‘two‐band’ pattern, which complicates band‐shape analysis using these traditional methods. In the past, the appearance of sub‐bands within or near the IVCT transition has been attributed to vibronic effects or localised d‐d transitions. Quantum‐chemical modelling of a series of rotational conformers of [ 1 ]+–[ 3 ]+ reveals the two components that contribute to the NIR absorption band envelope to be a π‐π* transition and an MLCT transition. The MLCT components only gain appreciable intensity when the orientation of the half‐sandwich ruthenium ligand spheres deviates from idealised cis (Ω P?Ru?Ru?P=0°) or trans (Ω P?Ru?Ru?P=180°) conformations. The increased steric demand of the supporting ligands, together with some underlying inter‐phosphine ligand T‐shaped CH???π stacking interactions across the series [ 1 ]+ to [ 2 ]+ to [ 3 ]+ results in local minima biased towards such non‐idealised conformations of the metal‐ligand fragments (Ω P?Ru?Ru?P=33–153°). Experimentally, this is indicated by appearance of multiple bands within the IR (C≡C) band envelopes and increasing intensity of the higher‐energy MLCT transition(s) relative to the π‐π* transition across the series, and the appearance of a pronounced ‘two‐band’ pattern in the experimental NIR absorption envelopes. These conformational effects and the methods of analysis presented here, which combine analysis of IR and NIR spectra with quantum‐chemical calculations on a range of energetically similar conformational minima, are expected to be quite general for mixed‐valence systems.  相似文献   

9.
The impact of the macrocyclic ligand on the electronic structure of two LNi? O2 biomimetic adducts, [Ni(12‐TMC)O2]+ (12‐TMC = 1,4,7,10‐tetramethyl‐1,4,7,10‐tetraazacyclododecane) and [Ni(14‐TMC)O2]+ (14‐TMC = 1,4,8,11‐tetramethyl‐1,4,8,11‐tetraazacyclotetradecane), has been inspected by means of difference‐dedicated configuration interaction calculations and a valence bond reading of the wavefunction. The system containing the 12‐membered macrocyclic ligand has been experimentally described as a side‐on nickel(III)‐peroxo complex, whereas the 14‐membered one has been characterized as an end‐on nickel(II)‐superoxide. Our results put in evidence the relationship between the steric effect of the macrocyclic ligand, the O2 coordination mode and the charge transfer extent between the Ni center and the O2 molecule. The 12‐membered macrocyclic ligand favors a side‐on coordination, a most efficient overlap between Ni 3d and O2 π* orbitals and, consequently, a larger charge transfer from LNi fragment to O2 molecule. The analysis of the ground‐state electronic structure shows an enhancement of the peroxide nature of the Ni? O2 interaction for [Ni(12‐TMC)O2]+, although a dominant superoxide character is found for both systems. © 2012 Wiley Periodicals, Inc.  相似文献   

10.
The Cholesky decomposition (CD) technique, used to approximate the two‐electron repulsion integrals (ERIs), is applied to the valence bond self‐consistent field (VBSCF) method. Test calculations on ethylene, C2nH2n+2, and C2nH4n?2 molecules (n = 1–7) show that the performance of the VBSCF method is much improved using the CD technique, and thus, the integral transformation from basis functions to VB orbitals is no longer the bottleneck in VBSCF calculations. The errors of the CD‐based ERIs and of the total energy are controlled by the CD threshold, for which a value of 10?6 ensures to control the total energy error within 10?6 Hartree. © 2016 Wiley Periodicals, Inc.  相似文献   

11.
The semiempirical method SCF MO LCAO in the CNDO/S valence approximation has been used to calculate tetramethyltetraselenafulvalene dimers (TMTSF)2. Parameters of the band structure of (TMTSF)2 +PF6 have been calculated with an accounting for charge transfer.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 28, No, 2, pp. 140–143, March–April, 1992.  相似文献   

12.
This study presents the experimental and theoretical study of highly internally Al‐doped TiO2 nanoparticles. Two synthesis methods were used and detailed characterization was performed. There were differences in the doping and the crystallinity, but the nanoparticles synthesized with the different methods share common features. Anatase to rutile transformation occurred at higher temperatures with Al doping. X‐ray photoelectron spectroscopy showed the generation of oxygen vacancies, which is an interesting feature in photocatalysis. In turn, the band‐gap energy and the valence band did not change appreciably. Periodic density functional calculations were performed to model the experimentally doped structures, the formation of the oxygen vacancies, and the band gap. Calculation of the density of states confirmed the experimental band‐gap energies. The theoretical results confirmed the presence of Ti4+ and Al3+. The charge density study and electron localization function analysis indicated that the inclusion of Al in the anatase structure resulted in a strengthening of the Ti?O bonds around the vacancy.  相似文献   

13.
14.
The photoelectrochemical behaviour of a Ru-doped TiO2 Crystal electrode of composition Ti0.97Ru0.03O2 in contact with aqueous electrolytes has been investigated. The substitution of Ru4+ for Ti4+ in the TiO2 lattice produces two main effects; (i) sensitization to visible light; (ii) reduction of the overpotential for O2 evolution, both in the dark and under illumination. Ru4+ eneryg levels constitute a narrow cationic band between the O2p valence band and the Ti3d conduction band, Ru4+ → Ti4+ electronic transitions being responsible for the subbandagap photoresponse. Besides Ru4+ ions at the semiconductor surface are easily oxidized under positive polarization of the electrode: the surface becomes charged positively and the Fermi level is pinned, which facilitates the transfer of charge from the filled levels of water molecules to the semiconductor conduction band, leading to O2 evolution. The transient photocurrent-time behaviour observed, both under bandgap and subbandgap illumination, is compared with that of undoped TiO2 and analyzed in terms of charge transfer at the semiconductor—electrolyte interface.  相似文献   

15.
Zn(meadtc)2(2,2′‐bipy) is a ZnS3N2 chromophore with a distorted square pyramidal geometry. The IR band at 1002 cm?1 and the bond valence sum value of 1.98 confirmed the monodentate dithiocarbamate in coordination. The non‐bonding Zn–S distance is 5.004(3) Å. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

16.
Photocatalytic hydrogen production is proposed as a sustainable energy source. Simultaneous reduction and oxidation of water is a complex multistep reaction with high overpotential. Photocatalytic processes involving semiconductors transfer electrons from the valence band to the conduction band. Sacrificial substrates that react with the photochemically formed holes in the valence band are often used to study the mechanism of H2 production, as they scavenge the holes and hinder charge carrier recombination (electron-hole pairs). Here, we show that the desired sacrificial agent is one forming a radical that is a fairly strong reducing agent, and whose oxidized form is not a good electron acceptor that might suppress the hydrogen evolution reaction (HER). In an acidic medium, methanol was found to fulfill both these requirements better than ethanol and propan-2-ol in the TiO2-(M0-NPs) (M=Au or Pt) system, whereas in an alkaline medium, the alcohols exhibit a reverse order of activity. Moreover, we report that CH2(OH)2 is by far the most efficient sacrificial agent in a nontrivial mechanism in acidic media. Our study provides general guidelines for choosing an appropriate sacrificial substrate and helps to explain the variance in the performance of alcohol scavenger-based photocatalytic systems.  相似文献   

17.
The electronic structure of CaB2O4(III) crystal obtained by using SIESTA program is reported in this article. It is observed that the band gap values are, respectively, 5.39 and 5.89 eV from our LDA and GGA calculations. The bond covalency and bond valence are calculated with a simplified method. For both Ca–O and B–O types of bond, the bond covalency has a decreasing trend with the increasing bond length. The result of bond covalency in explaining the interaction between atoms has been shown in good agreement with that of Mulliken population analysis. The ionic configuration for CaB2O4(III) in the fundamental state is estimated to be Ca+1.808B−0.68O−0.112. A summary of B–O distances for the four phases of CaB2O4 crystal from several works is also presented.  相似文献   

18.
The pyrimidine bridged binuclear complex (CN)5FepymRu(NH3)5- (I) was prepared in aqueous solution by mixing cquimolar of Fe(CN)5OH23? and Ru(NH3)5pym2+. Its mixed valence state molecule (CN)5FepymRu(NH3)5(II) was obtained upon oxidation of I by one equivalent of peroxydisulfate ion. Both binuclear complexes and corresponding Fe(II) and Ru(II) mononuclear complexes displayed a metal-to-ligand charge transfer absorption in 400–450 nm region. Rate constants of formation and dissociation of I and II were measured, and the values of kf (?103M?1s?1) and kd (?10?3-10?4 s?1) were consistent with kinetic results expected for the substitution of Fe(CN)5OH23? with di- and trivalent ligands. Cyclic voltammetry of I exhibited two one-electron steps of oxidation corresponding to [III, L, II] + e → [II, L, II] and [III, L, III] + e → [III, L, II], respectively. The mixed valence binuclear complex II showed an intervalence band at 955 nm with a molar extinction coefficient 5.80 × 102 M?1cm?1 and a half-width 5100 cm?l. The properties of the IT band conform to Hush's theory. Spectroscopic, electrochemical and kinetic results of II suggest that the mixed valence complex features a trapped - valence formulation with localized oxidation states of Fe(II) and Ru(III).  相似文献   

19.
The semiconductors aIBIIIc 2 Vi (AI = Cu, Ag; B111 = Ga, In; CVI = 5, Se) have found wide use in quantum electronics and nonlinear optics. Some of them show great promise as materials for solar batteries. The valence and vacant states of the components of AIBIIIC 2 VI were studied experimentally (SK X-ray emission and absorption spectra in sulfides) and theoretically using the cluster version of the local coherent potential. For this study, experimental literature data are used. The calculations were performed with inclusion of the d states of the noble metal lying at the top of the valence band and determining the nature of the chemical bond (p— d hybridization). Also included is the structural factor (anion shift from the equilibrium positions in a sphalerite structure), whose effect on the forbidden gap Eg is comparable to that of p— d hybridization. Lattice parameters a and c and anion shifts u are calculated in terms of Jaffe and Zunger theory using Pauling's tetrahedral radii. The calculated density of state curves agree well in shape and satisfactorily in energy with the corresponding X-ray and X-ray photoelectron spectra. Our results confirm that Bragg's idea on transferability and conservation of elementary bonds is applicable to the compounds under study. The estimated forbidden gaps Eg proved to be close to the experimental ones to an accuracy of approximately 0.2 eV for all compounds AIBIIIC 2 VI . Translated fromZhurnal Strukturnoi Khimii, Vol. 41, No. 3, pp. 515-524, May–June, 2000.  相似文献   

20.
The cluster (SrB2O4)2 existing in crystalline states is employed to model the electronic structure and linear optical properties of solid state compound SrB2O4. This compound is synthesized by high temperature solution reaction, and it crystallizes in the orthorhombic space group Pbcn with cell dimensions a = 1.1995(3), b = 0.4337(1), c = 0.6575(1) nm, V = 0.34202 nm3, and Z = 4, μ = 15.14 cm?1, Dcaled = 3.36 g/cm3. The dynamic refractive indices are obtained in terms of INDO/SCI following combination with the Sum‐Over‐States method. A width of the calculated gap is 4.424 eV between the valence band and conduction band, and the calculated average refractive index is 1.980 at a wavelength of 1.065 μm. The charge transfers from O2‐ anion orbitals to Sr2+cation orbitals make the significant contributions to linear polarizability in terms of analyses of atomic state density contributing to the valence and conduction bands.  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