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 accurate ground‐state potential energy surface of silicon dicarbide, SiC2, has been determined from ab initio calculations using the coupled‐cluster approach. Results obtained with the conventional and explicitly correlated coupled‐cluster methods were compared. The core‐electron correlation, higher‐order valence‐electron correlation, and scalar relativistic effects were taken into account. The potential energy barrier to the linear SiCC configuration was predicted to be 1782 cm?1. The vibration‐rotation energy levels of the SiC2, 29SiC2, 30SiC2, and SiC13C isotopologues were calculated using a variational method. The experimental vibration‐rotation energy levels of the main isotopologue were reproduced to high accuracy. In particular, the experimental energy levels of the highly anharmonic vibrational ν3 mode of SiC2 were reproduced to within 6.7 cm?1, up to as high as the v3 = 16 state.  相似文献   

2.
An accurate potential energy surface of sulfur dioxide, SO2, in its ground electronic state has been determined from ab initio calculations using the coupled‐cluster approach in conjunction with the correlation‐consistent basis sets up to septuple‐zeta quality. The results obtained with the conventional and explicitly correlated coupled‐cluster methods are compared. The role of the core–electron correlation, higher‐order valence–electron correlation, scalar relativistic, and adiabatic effects in determining the structure and dynamics of the SO2 molecule is discussed. The vibration‐rotation energy levels of the 32SO2 and 34SO2 isotopologues were predicted using a variational approach. It was shown that the inclusion of the aforementioned effects was mandatory to attain the “spectroscopic” accuracy. © 2017 Wiley Periodicals, Inc.  相似文献   

3.
The accurate ground‐state potential energy surface of germanium dicarbide, GeC2, has been determined from ab initio calculations using the coupled‐cluster approach. The core–electron correlation, higher‐order valence‐electron correlation, and scalar relativistic effects were taken into account. The potential energy surface of GeC2 was shown to be extraordinarily flat near the T‐shaped equilibrium configuration. The potential energy barrier to the linear CCGe configuration was predicted to be 1218 cm−1. The vibration–rotation energy levels of some GeC2 isotopologues were calculated using a variational method. The vibrational bending mode ν3 was found to be highly anharmonic, with the fundamental wavenumber being only 58 cm−1. Vibrational progressions due to this mode were predicted for the , , and states of GeC2. © 2018 Wiley Periodicals, Inc.  相似文献   

4.
5.
Following our investigations on hydrogen polyoxides, herein we employed coupled cluster theory in conjunction with Dunning's correlation consistent basis sets and density functional theory to study HOOOOOH (H2O5). The infrared spectra of H2O5 and its three deuterated isotopologues, as well as those of the five single‐substituted 18O isotopologues are discussed in detail. Internal valence coordinates were employed to classify the vibrational modes. The Raman activity is reported to help in the identification of hydrogen pentoxide. The suggested enthalpy of formation is ΔHf,298° (HOOOOH) = 1.4 ± 1.5 kcal/mol. This value includes corrections for relativistic and core‐valence effects as well as anharmonic corrections to Zero‐point energy corrections. © 2013 Wiley Periodicals, Inc.  相似文献   

6.
Two novel polymers exhibiting metal–organic frameworks (MOFs) have been synthesized by the combination of a metal ion with a benzene‐1,3,5‐tricarboxylate ligand (BTC) and 1,10‐phenanthroline (phen) under hydrothermal conditions. The first compound, poly[[(μ4‐benzene‐1,3,5‐tricarboxylato‐κ4O:O′:O′′:O′′′)(μ‐hydroxido‐κ2O:O)bis(1,10‐phenanthroline‐κ2N,N′)dizinc(II)] 0.32‐hydrate], {[Zn2(C9H3O6)(OH)(C12H8N2)2]·0.32H2O}n, denoted Zn–MOF, forms a two‐dimensional network in which a binuclear Zn2 cluster serves as a 3‐connecting node; the BTC trianion also acts as a 3‐connecting centre. The overall topology is that of a 63 net. The phen ligands serve as appendages to the network and interdigitate with phen ligands belonging to adjacent parallel sheets. The second compound, poly[[(μ6‐benzene‐1,3,5‐tricarboxylato‐κ7O1,O1′:O1:O3:O3′:O5:O5′)(μ3‐hydroxido‐κ2O:O:O)(1,10‐phenanthroline‐κ2N,N′)dimanganese(II)] 1.26‐hydrate], {[Mn2(C9H3O6)(OH)(C12H8N2)]·1.26H2O}n, denoted Mn–MOF, exists as a three‐dimensional network in which an Mn4 cluster serves as a 6‐connecting unit, while the BTC trianion again plays the role of a 3‐connecting centre. The overall topology is that of the rutile net. Phen ligands act as appendages to the network and form the `S‐shaped' packing mode.  相似文献   

7.
《Chemphyschem》2003,4(1):32-48
The coupled‐cluster approach, including single and double excitations and perturbative corrections for triple excitations, is capable of predicting molecular electronic energies and enthalpies of formation of small molecules in the gas phase with very high accuracy (specifically, with error bars less than 5 kJ mol?1), provided that the electronic wavefunction is dominated by the Hartree–Fock configuration. This capability is illustrated by calculations on molecules containing O–H and O–F bonds, namely OH, FO, H2O, HOF, and F2O. To achieve this very high accuracy, it is imperative to account for electron‐correlation effects in a quantitative manner, either by using explicitly correlated two‐particle basis functions (R12 functions) or by extrapolating to the limit of a complete basis. Besides taking into account harmonic zero‐point vibrational energies, it is also necessary to account for anharmonic corrections to the zero‐point vibrational energies, to include the core orbitals into the coupled‐cluster calculations, and to account for spin–orbit corrections and scalar relativistic effects. These additional corrections constitute small but significant contributions in the range of 1–4 kJ mol?1 to the enthalpies of formation of the aforementioned molecules. The highly accurate coupled‐cluster results, obtained by employing R12 functions and by including various corrections, are compared with standard Kohn–Sham density‐functional calculations as well as with the Gaussian‐2 and complete‐basis‐set model chemistries.  相似文献   

8.
Oxygen‐rich scandium cluster anions ScO3–5? are prepared by laser ablation and allowed to react with n‐butane in a fast‐flow reactor. A time‐of‐flight mass spectrometer is used to detect the cluster distribution before and after the reactions. The ScO3? and ScO4? clusters can react with n‐butane to produce ScO3H?, ScO3H2?, and ScO4H?, while the more oxygen‐rich cluster ScO5? is inert. The experiment suggests that unreactive cluster isomers of ScO3? and ScO4? are also present in the cluster source. Density functional theory and ab initio methods are used to calculate the structures and reaction mechanisms of the clusters. The theoretical results indicate that the unreactive and reactive cluster isomers of ScO3,4? contain peroxides (O22?) and oxygen‐centered radicals (O.?), respectively. The mechanisms and energetics for conversion of unreactive O22? to reactive O.? species are also theoretically studied.  相似文献   

9.
A new four‐dimensional intermolecular potential energy surface for CS2 dimer is obtained by ab initio calculation of the interaction energies for a range of configurations and center‐of‐mass separation distances for the first time. The calculations were performed using the supermolecular approach at the Møller–Plesset second‐order perturbation (MP2) level of theory with the augmented correlation consistent basis sets (aug‐cc‐pVxZ, x = D, T) and corrected for the basis‐set superposition error using the full counterpoise correction method. A two‐point extrapolation method was used to extrapolate the calculated energy points to the complete basis set limit. The effect of using the higher levels of theory, quadratic configuration interaction containing single, double, and perturbative triple excitations QCISD(T) and coupled cluster singles, doubles and perturbative triples excitations CCSD(T), on the shape of potential energy surface was investigated. It is shown that the MP2 level of theory apparently performs extremely poorly for describing the intermolecular potential energy surface, overestimating the total energy by a factor of nearly 1.73 in comparison with the QCISD(T) and CCSD(T) values. The value of isotropic dipole–dipole dispersion coefficient (C6) of CS2 fluid was obtained from the extrapolated MP2 potential energy surface. The MP2 extrapolated energy points were fitted to well‐known analytical potential functions using two different methods to represent the potential energy surface analytically. The most stable configuration of the dimer was determined at R = 6.23 au, α = 90°, β = 90°, and γ = 90°, with a well depth of 3.980 kcal mol?1 at the MP2 level of theory. Finally, the calculated second virial coefficients were compared with experimental values to test the quality of the presented potential energy surface. © 2010 Wiley Periodicals, Inc. J Comput Chem, 2011.  相似文献   

10.
In the structure of trans‐bis(ethanol‐κO)tetrakis(1H‐imidazole‐κN3)copper(II) bis[μ‐N‐(2‐oxidobenzylidene)‐D,L‐glutamato]‐κ4O1,N,O2′:O2′4O2′:O1,N,O2′‐bis[(1H‐imidazole‐κN3)cuprate(II)], [Cu(C3H4N2)4(C2H6O)2][Cu2(C15H14N3O5)2], both ions are located on centres of inversion. The cation is mononuclear, showing a distorted octahedral coordination, while the anion is a binuclear centrosymmetric dimer with a square‐pyramidal copper(II) coordination. An extensive three‐dimensional hydrogen‐bonding network is formed between the ions. According to B3LYP/6–31G* calculations, the two equivalent components of the anion are in doublet states (spin density located mostly on CuII ions) and are coupled as a triplet, with only marginal preference over an open‐shell singlet.  相似文献   

11.
(S)‐1‐(Methylaminocarbonyl)‐3‐phenylpropanaminium chloride (S2·HCl), C10H15N2O+·Cl, crystallizes in the orthorhombic space group P212121 with a single formula unit per asymmetric unit. (5R/S)‐5‐Benzyl‐2,2,3‐trimethyl‐4‐oxoimidazolidin‐1‐ium chloride (R3 and S3), C13H19N2O+·Cl, crystallize in the same space group as S2·HCl but contain three symmetry‐independent formula units. (R/S)‐5‐Benzyl‐2,2,3‐trimethyl‐4‐oxoimidazolidin‐1‐ium chloride monohydrate (R4 and S4), C13H19N2O+·Cl·H2O, crystallize in the space group P21 with a single formula unit per asymmetric unit. Calculations at the B3LYP/6–31G(d,p) and B3LYP/6–311G(d,p) levels of the conformational energies of the cation in R3, S3, R4 and S4 indicate that the ideal gas‐phase global energy minimum conformation is not observed in the solid state. Rather, the effects of hydrogen‐bonding and van der Waals interactions in the crystal structure cause the molecules to adopt higher‐energy conformations, which correspond to local minima in the molecular potential energy surface.  相似文献   

12.
The determination of minima and saddle points on the potential energy surfaces of the hydrogen bonded species O2?HF and O2?H2O is performed with unrestricted Hartree-Fock calculations. Geometries, electron density distributions, and relative energies for every stationary point are reported. Only one true minimum is found for O2?HF and for O2?H2O, and this approximately corresponds to a structure where the partially positive hydrogen atom is located along one of the superoxide ion electron lone-pair directions. Calculated ΔH, ΔS, and ΔG values for the reaction between O2? and H2O are in good agreement with experimental data.  相似文献   

13.
The reaction of 0.67 molar equivalents of the O,N,O′‐tridentate zwitterionic Schiff base (2Z,4E)‐4‐[(2‐hydroxyphenyl)iminio]pent‐2‐en‐2‐olate (H2L) with one equivalent of zinc(II) acetate in methanol affords a novel trinuclear ZnII cluster, di‐μ‐acetato‐1:2κ2O:O′;2:3κ2O:O′‐dimethanol‐1κO,3κO‐bis{μ‐2‐[(2E,3Z)‐4‐oxidopent‐3‐en‐2‐ylideneamino]phenolato}‐1:2κ4O2,N,O4:O4;2:3κ4O4:O2,N,O4‐trizinc(II), [Zn3(C11H11NO2)2(C2H3O2)2(CH4O)2], (I), in which two bridging acetate ligands link the terminal square‐based pyramidal ZnII ions to the approximately tetrahedral ZnII ion at the core of the cluster. The ZnO4 coordination group of the central ZnII ion is established by two bridging phenolate and two bridging acetate O atoms. The remaining four coordination sites of each terminal ZnII ion are occupied by methanol and deprotonated H2L. Furthermore, the Zn‐bound methanol hydroxyl groups are involved in complementary hydrogen bonding with the Zn‐bound enolate O atom of a neighbouring molecule, about an inversion centre in each case. The structure of (I) is therefore best described as an extended one‐dimensional hydrogen‐bonded chain of trinuclear ZnII clusters.  相似文献   

14.
Schiff bases are considered `versatile ligands' in coordination chemistry. The design of polynuclear complexes has become of interest due to their facile preparations and varied synthetic, structural and magnetic properties. The reaction of the `ligand complex' [CuL] {H2L is 2,2′‐[propane‐1,3‐diylbis(nitrilomethanylylidene)]diphenol} with Ni(OAc)2·4H2O (OAc is acetate) in the presence of dicyanamide (dca) leads to the formation of bis(dicyanamido‐1κN1)bis(dimethyl sulfoxide)‐2κO,3κO‐bis{μ‐2,2′‐[propane‐1,3‐diylbis(nitrilomethanylylidene)]diphenolato}‐1:2κ6O,O′:O,N,N′,O′;1:3κ6O,O′:O,N,N′,O′‐dicopper(II)nickel(II), [Cu2Ni(C17H16N2O2)2(C2N3)2(C2H6OS)2]. The complex shows strong absorption bands in the frequency region 2155–2269 cm−1, which clearly proves the presence of terminal bonding dca groups. A single‐crystal X‐ray study revealed that two [CuL] units coordinate to an NiII atom through the phenolate O atoms, with double phenolate bridges between CuII and NiII atoms. Two terminal dca groups complete the distorted octahedral geometry around the central NiII atom. According to differential thermal analysis–thermogravimetric analysis (DTA–TGA), the title complex is stable up to 423 K and thermal decomposition starts with the release of two coordinated dimethyl sulfoxide molecules. Free H2L exhibits photoluminescence properties originating from intraligand (π–π*) transitions and fluorescence quenching is observed on complexation of H2L with CuII.  相似文献   

15.
An in‐depth spectroscopic EPR investigation of a key intermediate, formally notated as [PVIVVVMo10O40]6? and formed in known electron‐transfer and electron‐transfer/oxygen‐transfer reactions catalyzed by H5PV2Mo10O40, has been carried out. Pulsed EPR spectroscopy have been utilized: specifically, W‐band electron–electron double resonance (ELDOR)‐detected NMR and two‐dimensional (2D) hyperfine sub‐level correlation (HYSCORE) measurements, which resolved 95Mo and 17O hyperfine interactions, and electron–nuclear double resonance (ENDOR), which gave the weak 51V and 31P interactions. In this way, two paramagnetic species related to [PVIVVVMo10O40]6? were identified. The first species (30–35 %) has a vanadyl (VO2+)‐like EPR spectrum and is not situated within the polyoxometalate cluster. Here the VO2+ was suggested to be supported on the Keggin cluster and can be represented as an ion pair, [PVVMo10O39]8?[VIVO2+]. This species originates from the parent H5PV2Mo10O40 in which the vanadium atoms are nearest neighbors and it is suggested that this isomer is more likely to be reactive in electron‐transfer/oxygen‐transfer reaction oxidation reactions. In the second (70–65 %) species, the VIV remains embedded within the polyoxometalate framework and originates from reduction of distal H5PV2Mo10O40 isomers to yield an intact cluster, [PVIVVVMo10O40]6?.  相似文献   

16.
The croconate dianion is a highly versatile ligand with two tautomeric forms making it useful for building large superstructures in the solid state. The single‐crystal X‐ray structures of PbII– and CdII–croconate coordination polymers, namely catena‐poly[[[diaqualead(II)]‐μ‐croconato‐κ4O1,O2:O3,O4] monohydrate], {[Pb(C5O5)(H2O)2]·H2O}n, 1 , and catena‐poly[[triaquacadmium(II)]‐μ‐croconato‐κ4O1,O2:O3,O4], [Cd(C5O5)(H2O)3]n, 2 , have been determined. Both polymers form one‐dimensional (1D) structures; 1 is a nonplanar 1D zigzag coordination polymer extended along the crystallographic b axis, whereas 2 is a planar 1D ribbon parallel to the [101] direction. In 2 , three H2O molecules are coordinated directly to the metal atom, while in 1 , only two H2O molecules are directly coordinated to the metal atom. A third interstitial H2O molecule is involved in hydrogen bonding with O atoms of the croconate ligands of an adjacent layer and other H2O molecules, resulting in stacked double layers parallel to the [105] plane. Solid‐state FT–IR and solution UV–Vis spectra also substantiate the croconate coordination.  相似文献   

17.
We studied the attraction between [C2Hn] and Tl(I) in the hypothetical [C2Hn–Tl]+ complexes (n = 2,4) using ab initio methodology. We found that the changes around the equilibrium distance C–Tl and in the interaction energies are sensitive to the electron correlation potential. We evaluated these effects using several levels of theory, including Hartree–Fock (HF), second‐order Møller–Plesset (MP2), MP4, coupled cluster singles and doubles CCSD(T), and local density approximation augmented by nonlocal corrections for exchange and correlation due to Becke and Perdew (LDA/BP). The obtained interaction energies differences at the equilibrium distance Re (C–Tl) range from 33 and 46 kJ/mol at the different levels used. These results indicate that the interaction between olefinic systems and Tl(I) are a real minimum on the potential energy surfaces (PES). We can predict that these new complexes are viable for synthesizing. At long distances, the behavior of the [C2Hn]–Tl+ interaction may be related mainly to charge‐induced dipole and dispersion terms, both involving the individual properties of the olefinic π‐system and thallium ion. However, the charge‐induced dipole term (R?4) is found as the principal contribution in the stability at long and short distances. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

18.
A simple and effective synthetic route to homo‐ and heteroleptic rare‐earth (Ln = Y, La and Nd) complexes with a tridentate Schiff base anion has been demonstrated using exchange reactions of rare‐earth chlorides with in‐situ‐generated sodium (E)‐2‐{[(2‐methoxyphenyl)imino]methyl}phenoxide in different molar ratios in absolute methanol. Five crystal structures have been determined and studied, namely tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐κ3O1,N,O2)lanthanum, [La(C14H12NO2)3], ( 1 ), tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐κ3O1,N,O2)neodymium tetrahydrofuran disolvate, [La(C14H12NO2)3]·2C4H8O, ( 2 )·2THF, tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato)‐κ3O1,N,O23O1,N,O22N,O1‐yttrium, [Y(C14H12NO2)3], ( 3 ), dichlorido‐1κCl,2κCl‐μ‐methanolato‐1:2κ2O:O‐methanol‐2κO‐(μ‐2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐1κ3O1,N,O2:2κO1)bis(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato)‐1κ3O1,N,O2;2κ3O1,N,O2‐diyttrium–tetrahydrofuran–methanol (1/1/1), [Y2(C14H12NO2)3(CH3O)Cl2(CH4O)]·CH4O·C4H8O, ( 4 )·MeOH·THF, and bis(μ‐2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐1κ3O1,N,O2:2κO1)bis(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐2κ3O1,N,O2)sodiumyttrium chloroform disolvate, [NaY(C14H12NO2)4]·2CHCl3, ( 5 )·2CHCl3. Structural peculiarities of homoleptic tris(iminophenoxide)s ( 1 )–( 3 ), binuclear tris(iminophenoxide) ( 4 ) and homoleptic ate tetrakis(iminophenoxide) ( 5 ) are discussed. The nonflat Schiff base ligand displays μ2‐κ3O1,N,O2O1 bridging, and κ3O1,N,O2 and κ2N,O1 terminal coordination modes, depending on steric congestion, which in turn depends on the ionic radii of the rare‐earth metals and the number of coordinated ligands. It has been demonstrated that interligand dihedral angles of the phenoxide ligand are convenient for comparing steric hindrance in complexes. ( 4 )·MeOH has a flat Y2O2 rhomboid core and exhibits both inter‐ and intramolecular MeO—H…Cl hydrogen bonding. Catalytic systems based on complexes ( 1 )–( 3 ) and ( 5 ) have demonstrated medium catalytic performance in acrylonitrile polymerization, providing polyacrylonitrile samples with narrow polydispersity.  相似文献   

19.
The title compound, poly[aqua(μ2‐1H‐imidazole‐4‐carboxylato‐κ3N3,O:O′)hemi(μ2‐oxalato‐κ4O1,O2:O1′,O2′)cadmium(II)], [Cd(C4H3N2O2)(C2O4)0.5(H2O)]n, exhibits a two‐dimensional network. The CdII cation is coordinated to one N atom and two carboxylate O atoms from two 1H‐imidazole‐4‐carboxylate (Himc) ligands, two carboxylate O atoms from the bridging oxalate anion and one ligated water molecule; these six donor atoms form a distorted octahedral configuration. The oxalate anion lies on a centre of inversion. The Himc ligands connect the CdII cations to form –Cd–Himc–Cd–Himc–Cd– zigzag chains, with a Cd...Cd separation of 5.8206 (6) Å along the b direction, which are further linked by tetradentate oxalate anions to generate a two‐dimensional herringbone architecture in the ab plane. These layers are extended to form a three‐dimensional supramolecular framework via O—H...O and N—H...O hydrogen bonds and π–π stacking interactions. The solid‐state photoluminscent behaviour of the title compound has been investigated at room temperature.  相似文献   

20.
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