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 equilibrium structures and relative stabilities of BN-doped fullerenes C70−2x(BN)x (x=1–3) have been studied at the AM1 and MNDO level. The most stable isomers of C70−2x(BN)x have been found out and their electronic properties have been predicted. The calculation results show that the BN substituted fullerenes C70−2x(BN)x have considerable stabilities, though they are less stable than their all carbon analog. For C68BN, the isomers whose BN is located in the most chemically active bonds of C70 (namely B and A) are among the most stable species, of which B is predicted to be the ground state. The stabilities of C68BN decrease and the dipole moments increase with increasing the distance between the heteroatoms. For C66(BN)2, the lowest energy species is the isomer in which the B–N–B–N bond is formed; For C64(BN)3, the most stable species should have three BN units located in the same hexagon to form B–N–B–N–B–N ring. The ionization potentials and the affinity energies of the most stable species of BN-doped C70 are almost the same as those of C70 because of the isoelectronic relationship. The ionization potentials and affinity energies depend on the relative position of the heteroatoms in C68BN, the chemical reactivities of the isomers whose heteroatoms are well separated should differ significantly from their all carbon analog.  相似文献   

2.
For a closed-shell MO configuration with 2n electrons which occupy n non-degenerate canonical MOs, it is deduced that the RHF energy, Σni=1[2H0nnj-1(2Jij-Kij)], may be expressed in Hückel-like form as 2Σni-1ε, −Σni-1[ji(λ+1)+1,(λ+2)] with λ=2(n-i). The li(λ+1) and Ii(λ+2) are the ionization potentials for the HOMO ψ, which arises after λ and λ+1 electrons have been successively removed from the initial configuration.  相似文献   

3.
The structure of cyclopentadienyl(duroquinone)cobalt dihydrate, (C5H5)Co-[(CH3)4C6O2]·2H2O, has been determined by three-dimensional X-ray analysis. The crystal structure consists of discrete cyclopentadienyl(duroquinone)cobalt molecules linked together by a complex network of hydrogen bonds between water molecules and duroquinone oxygen atoms. Each (C5H5)Co[(CH3)4C6O2] molecule consists of a cobalt atom sandwiched between a cyclopentadienyl ring and a duroquinone ring. A detailed comparison of the molecular parameters of this complex with those of closely related complexes is given. Crystallographic evidence that the metal---duroquinone interaction in cyclopentadienyl(duroquinone)cobalt dihydrate is considerably stronger than that in the electronically-equivalent 1,5-cyclooctadiene(duroquinone)nickel complex is given not only by the metal---C(olefin) distances being 0.12 Å (av) shorter in the duroquinone---cobalt complex [viz., 2.104(8) Å vs. 2.222(7) Å] but also by the much greater C2v-type distortion of the duroquinone ring from the planar D2h configuration in free duroquinone. The compound crystallizes with two formula species in a triclinic unit cell of symmetry P and reduced cell dimensions á = 8.60 Å, b = 9.00 Å, c = 10.15 Å, = 87° 34′, β = 84° 10′, γ = 73° 44′. Least-squares refinement yielded final unweighted and weighted discrepancy factors of R1 = 10.8% and R2 = 12.0%, respectively, for 2481 independent diffraction maxima collected photographically.  相似文献   

4.
Activities of solutes and compositions of solutions may be expressed corretly in terms of molarity (c), molality (m) or mole fraction (x), leading to corresponding equilibrium constants Kc, Km, or Kx. Equations for differences between ΔG°c, ΔG°m, and ΔG°x values are derived. Common errors in calculations involving (dlnKc/dT) and (dlnKc/dP) are identified and remedies for these errors are presented.  相似文献   

5.
The gas-phase stabilities of cluster ions SF+m (SF6)n with m = 0−5 were determined by using a high pressure mass spectrometer. The bond energies of SF+m (SF6)1 were found to be less than 10 kcal/mol and to decrease with m = 0 → 5. There appear to be rather large gaps in the bond energies between n = 1 and 2 for the clusters SF+m (SF6)n with m = 0−4. The structures of SF+5, SF+ (SF6)1, SF+3 (SF6)1, and SF+5 (SF6)1 were investigated by ab initio molecular orbital calculations. For SF+5, the D3h geometry is found to be most stable andC4v is a transition state of the Berry pseudorotation. For the ion-molecule complexes, the “on-top hat” models were found to be the most stable structures.  相似文献   

6.
Ab initio calculations are carried out at UB3LYP/6-311++G (3df, 2p) levels of theory, on electrocyclic thermal cleavage of four (S) derivatives of diaziridines, 1X-R, to their corresponding (Z) and (E) azomethine imides (2X-Z, 2X-E, 3X-Z and 3X-E), where X=–H, –Me, t-Bu and Ph. Cleavage of 1X-R Series to 2X-Z (Path 1) emerged as the more favored, for producing the most stable products, 2X-Z. In IRC calculations that were shown in Paths 1 and 2, C6–N1 bond was cleavage, before reaching reaction rate determinating step (transition state).  相似文献   

7.
Photodecomposition of 10 different molybdenum and tungsten mixed carbonyl complexes, [M(CO)3(B)(A)]I2 where B=o-phenanthroline or bipyridyl, A=3-(2-propynyl)thio-4,5-diphenyl-4H-1,2,4-triazole (TRZA) or S-propynyl-2-thio benz-imidazole (BIMDA) and 2(2-propynyl-thio(5-phenyl)-1,3,4-oxadiazole (OXA). M(CO)3(TRZA)I2, [M(CO)2(PPh3)X(TRZA)IY]IZ where M=Mo, X, Y and Z=1 and M=W, X and Z=2, Y=0, have been performed at 365 nm in oxygen saturated chloroform at 25 °C. The absorbance spectrum of these complexes have been recorded with the time of irradiation in order to examine the kinetics of photodecay.

The apparent rate constant (Kd) for the first-order reaction have been calculated and found to be (3.32–4.79)×10−5 s−1. The primary quantum yields (Φ) has also been calculated and were in the range (8.33–12.1)×10−4. The mechanism of the photodecomposition has been suggested according to the kinetic, and photoproduct analysis data, and is similar to reaction of photo-oxidative degradation of polluted molecules in the water.  相似文献   


8.
Ab initio molecular orbital theory was used to determine the equilibrium structure and vibrational frequencies of Fe2Cl6 and FeAlCl6. The equilibrium structure the Fe2Cl6 dimer has D2h symmetry with a planar arrangement of the four membered {FeClbrFeClbr} ring, similar to the Al2Cl6 dimer. The calculated bond distances and vibrational frequencies are in good agreement with experiment. The potential energy surface for the puckering of the {FeClbrFeClbr} ring is extremely flat. This prevents an unambiguous assignment of either D2h or C2v symmetry to the Fe2Cl6 structure in electron diffraction measurements. The FeAlCl6 molecule is found to have a C2v structure similar to Fe2Cl6 with vibrational frequencies in good agreement with experiment.  相似文献   

9.
Some aspects of the theory of LASIN (laser assisted surface ion neutralization) are discussed, with emphasis on the physical origins of the so-called double-peak structures found in some calculations of the charge-transfer (neutralization) probability, P, as a function of the laser frequency η. These two peaks have been called the first peak at η ≈ ηm = Om (in a.u.), where o(m) is the electronic energy level of the ion/atom (middle of the solid's valence band) and the second peak, a much larger peak at η ≈ 1.3 ηm, respectively.

We show that these double-peak structures are all special cases of multiple-peak structures which result from quantum interference effects, and that, in fact, the second peak is to be regarded as the main resonance peak. This result is interesting in itself, because it is the first peak which has heretofore been considered the main resonance peak.

To simplify the discussion, a two-level model is adapted, which represents the solid valence band by a single level at m. Clarification of the physical reason for the multiple peaks is based on the semiclassical theory of nonadiabatic transitions, in which the peaks are due to the phase difference between the two adiabatic paths that arise from the diagonalization of the two-level hamiltonian.

With the electronic hopping potential modelled by V(t) = Vosech(λt), and the laser potential by W(t) = Wosech(λt) cos(πt + δ), in the usual notation, an approximate analytical expression for P(η) is presented for the case Wo/Vo < 1, which covers most of the previous treatments, and is in good agreement with the exact results.  相似文献   


10.
The CCSD(T)/11e-RECP//MP2/11e-RECP method was used to explore the potential energy surfaces (PESs) of the formation of Agn (n = 2–6) clusters. Two kinds of reaction mechanisms were revealed in the formation of Agn clusters, the association mechanism for the formation of Ag2, Ag5, and Ag6 clusters and the association–isomerization mechanism for the formation of Ag3 and Ag4 clusters. Based on the canonical transition state theory, the calculated rate constants of the formation of Agn clusters displayed an odd–even effect: the rate constants of formation of Agn clusters with odd number were larger than those with even number. The rate constant of formation of Ag4 was the lowest, whereas that of Ag5 was the highest among Agn (n = 2–6) clusters. The formation of Ag4 was the most difficult step in the aggregation process of the silver clusters. The formation of Ag4 may be related with the critical point in the silver aggregation process.  相似文献   

11.
12.
CaRgn+ (Rg=He, Ne, Ar) complexes with n=1–4, are investigated by performing using the B3LYP/6-311+G (3df) density functional theory calculations. The CaHen+ (n=1–4) complexes are found to be stable. In the case of CaNen+ and CaArn+, stable structures and stationary point were found only for n=1 and 2. For n=3 in the C3V and the D3h point group as well as for n=4 in the Td (tetrahedral) point group a saddle point (imaginary frequency) is observed and global minimum could be obtained along the potential energy surface.  相似文献   

13.
A Doppler-based velocity selection technique has been used to measure the relative velocity dependence of the cross sections σji,Δr) for rotationally inelastic collisions from level ji to ji + Δν1 = 8,22,42) in 7Li*2 A 1Σ+u)—Xe. The σjν±2r) are strongly attenuated at a smaller νr by “torque averaging” due to molecular rotation; in contrast, for large |Δ|, σj = νrn (1 n 2). An empirical intermolecular potential which reproduces these types of behavior for 3-D classical trajectories is exhibited.  相似文献   

14.
The synthesis and reactivity of {(η5-C5H4SiMe3)2Ti(CCSiMe3)2} MCl2 (M = Fe: 3a; M = Co: 3b; M = Ni: 3c) is described. The complexes 3 are accessible by the reaction of (η5-C5H4SiMe3) 2Ti(CSiMe3)2 (1) with equimolar amounts of MCl2 (2) (M = Fe, Co, Ni). 3a reacts with the organic chelat ligands 2,2′-dipyridyl (dipy) (4a) or 1,10-phenanthroline (phen) (4b) in THF at 25°C to afford in quantitative yields (η5-C5H4SiMe3)2Ti(CSiMe3)2 (1) and [Fe(dipy)2]Cl2 (5a) or [Fe(phen)2]Cl2 (5b). 1/n[CuIHal]n (6) or 1/n[AgIHal]n (7) (Hal = Cl, Br) react with {(η5 -C5H4SiMe3)2Ti(CCSiMe3)2}FeCl2 (3a), by replacement of the FeCl2 building block in 3a, to yield the compounds {(η5-C5H4SiMe3)2Ti(C CSiMe3)2}CuIHal (8) or {(η5-C5H4SiMe3)2Ti(CSiMe3)2}AgIHal (9) (Hal = Cl, Br), respectively. In 8 and 9 each of the two Me3SiCC-units is η2-coordinated to monomeric CuI Hal or AgIHal moieties. Compounds 8 and 9 can also be synthesized by the reaction of (η5-C5H4SiMe3)2 Ti(CSiMe3)2 (1) with 1/n[CuIHal]n (6) or 1/n [AgIHal]n (7) in excellent yields. All new compounds have been characterized by analytical and spectroscopic data (IR, 1H-NMR, MS). The magnetic moments of compounds 3 were measured.  相似文献   

15.
The syntheses of Bromodisilanes BrnSi2H6−n and Iododisilanes InSi2H6−n (n = 1, 2, 3, 4, 5), starting from caryldisilanes ArnSi2H6−n (Ar = phenyl, -naphthyl, mesityl) are reported. The 29Si-NMR-spectra of all compounds, including 29Si29Si-coupling constants, have been measured.

Zusammenfassung

Ausgehend von Aryldisilanen ArS2H6−n, (Ar = Phenyl, -Naphthyl, Mesityl) wurden die Bromdisilane BrnSi2H6−n, und Ioddisilane InSi2H6−n, (n = 1, 2, 3, 4, 5) synthetisiert. Die 29Si-NMR-Spektren aller Verbindungen, (eingeschlossen 29Si29Si-Kopplungskonstanten) wurden vermessen.  相似文献   


16.
139La-NMR chemical shifts were measured for several anionic complexes of formulae Li(C4H8O2)3/2 [La(ν3-C3H5)4], [Li(C4H8O2)2][Cp′nLa(ν3-C3]H5)4−n] (Cp′ = Cp(ν5-C5H5); n = 1, 2 and Cp′ = Cp * (ν5-C5Me5); N = 1) and Li[RnLa(ν3-C3H4)4n] (R = N(SiMe3)2; n = 1, 2 and R = CCsIMe3; n = 4), as well as for neutral compounds for formulae La(ν3-C3H5)3Ln (L = (C4H8O2)1.5, (HMPT)2, TMED), Cp′nLa(ν3-C3H5)3−n (Cp′= Cp(ν5-Cp5H5), Cp *(ν5-C5Me5); n = 1, 2) and La(ν3-C3H2)2X(THF)2 X = Cl, Br, I). Typical ranges of the 139La-NMR chemical shifts were found for the different types of complex independent of number and kind of organyl groups directly bonded to lanthanum.

Zusammenfassung

139La-NMR-Spektroskopie wurde an einer Reihe anionischer Allyllanthanat(III)-Komplexe der Zusammensetzung ]- [La)ν3-C3H5)4, [Li(C4H8)2][Cp′nLa(ν3-C3H5)4−n(Cp′ = Cp(ν5-C5H5); n = 1, 2 und Cp′ = Cp * (ν5-C5Me5); N = 1) und Li[RnLa(ν3-C3H5)4−n (R = B(SiMe3)2; n = 1, 2 und R = CCSiMe3; n = 4 sowie neutraler Allyllanthan(III)-Komplexe der Zusammensetzung La(ν3-C3H5)3Ln (Ln = (C4H8O2)1.5, (HMPT)2, TMED), Cp′n, La(ν3-C3H5)3−n (Cp′ = Cp(ν5-C5H5), Cp * (ν5- Cp5Me5); n = 1, 2) und La(ν3-Cp3H5)2X(THF)2 (X = Cl, Br, I) durchgefürt. In Abhängikeit von der Anzahl und der Art der am Lanthan gebundenen Gruppen wurden für die verschieden Komplextypen charakteristische Resonanzbereiche ermittelt.  相似文献   


17.
Ferrocenyl-1,2-diketones FcCOCOR, 3, [Fc = (C5H5)Fe(C5H4)] can be prepared by oxidation of acylferrocenes FcCOCH2R or, more efficiently, by oxidation of the isomeric ketones FcCH2COR, 2. The ketones 2 are in turn readily synthesized from the salt (FcCH2PPh3)+I via the acylated salts [FcCH(COR)PPh3]+I. The haloacylferocenes FcCOCClx H3−x (x = 1, 2, 3, of which the x = 2 example is synthetically equivalent to a diketone) are synthesized by Friedel—Crafts acylation of ferrocene using CClxH3−xCOCl/AlCl3, but the reaction proceeds via two parallel pathways, one giving the normal acyl derivatives FcCOCClxH3−x and the other giving the reduced products FcCOCClx−1H4−x. Two diketones FcCOCOFc 3b and FcCOCOC6H4Ph 3c have been structurally characterised by single-crystal X-ray diffraction.  相似文献   

18.
A series of CexPr1−xO2−δ mixed oxides were synthesized by a sol–gel method and characterized by Raman, XRD and TPR techniques. The oxidation activity for CO, CH3OH and CH4 on these mixed oxides was investigated. When the value x was changed from 1.0 to 0.8, only a cubic phase CeO2 was observed. The samples were greatly crystallized in the range of the value x from 0.99 to 0.80, which is due to the formation of solid solutions caused by the complete insertion of Pr into the CeO2 crystal lattices. Raman bands at 465 and 1150 cm−1 in CexPr1−xO2−δ samples are attributed to the Raman active F2g mode of CeO2. The broad band at around 570 cm−1 in the region of 0.3 ≤ x ≤ 0.99 can be linked to oxygen vacancies. The new band at 195 cm−1 may be ascribed to the asymmetric vibration caused by the formation of oxygen vacancies. The TPR profile of Pr6O11 shows two reduction peaks and the reduction process is followed: . The reduction temperature of CexPr1−xO2−δ mixed oxides is lower than those of Pr6O11 or CeO2. TPR results indicate that CexPr1−xO2−δ mixed oxides have higher redox properties because of the formation of CexPr1−xO2−δ solid solutions. The presence of the oxygen vacancies favors CO and CH3OH oxidation, while the activity of CH4 oxidation is mostly related to reduction temperatures and redox properties.  相似文献   

19.
The electron scattering pattern of gaseous dicyclopentadienylberyllium, Cp2Be, has been recorded from s = 2.00 to 39.00 Å−1 with a nozzle temperature of about 120°C. Molecular models of D5d symmetry or models containing one π-bonded and one σ-bonded Cp ring are not compatible with the data. The possibility the gaseous Cp2Be consists fo a mixture D5d and π-Cp, σ-Cp conformers is considered and rejected. A model of C5v symmetry can be brought into satisfactory agreement with the data. It is also found that a slip sandwich model obtained from the C5v model by moving sideways the ring which is at the greatest distance from Be, while keeping the two rings essentially parallel is compatible with the electron diffraction data. The best fit between experimental and calculated intensity curves is obtained with a model with a sideways slip of 0.8(1) Å. This model is similar to that indicated by the X-ray diffraction investigations by Wong and coworkers [4,5]. It is suggested that the potential energy of the molecule does not change much as the magnitude of the slip changes and that the molecule thus undergoes large amplitude vibration.  相似文献   

20.
The structure of the complex [Ni(hmt)(NCS)2(H2O)2]n, assembled by hexamethylenetetramine (hmt) and octahedral Ni(II), is reported. Crystal data: Fw 351.07, a=9.885(10) Å, b=12.06(1) Å, c=12.505(8) Å, β=114.41(4)°, V=1357(1) Å3, Z=4, space group=C2/c, T=173 K, λ(MoK)=0.71070 Å, ρcalc=1.718 gcm−1, μ=17.44 cm−1, R=0.099, Rw=0.145. The tetrahedral assembling template effect of the hmt molecule is completed by two coordination bonds and two hydrogen interactions. The UV–vis absorption spectrum of this complex [Ni(hmt)(NCS)2(H2O)2]n with a two-dimensional network is determined in the range of 5000–35000 cm−1 at room temperature. The observed spectrum is discussed and explained perfectly by the scaling radial theory proposed by us. The two-dimensional structure has no apparent effects on the d–d transitions of the central Ni(II) ion. The IR spectrum and the GT curve of the complex were also measured and clearly reflect its structural properties.  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