首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到18条相似文献,搜索用时 265 毫秒
1.
采用ab initio HF, MP2方法和密度泛函理论方法, 对具有D2h和D4d构型的膦配体稳定的过渡金属团簇[Au@Au8(PR3)8]3+(R=Me, OMe, H, F, Cl, CN)进行了几何结构、 电子结构及团簇稳定性等方面的研究. 计算表明, 与D2h构型相比, D4d构型更稳定, 两者能量相差约5~10 kJ/mol. SVWN局域泛函能够对团簇的几何结构给予较准确的描述, MP2方法对团簇的结构参数有所低估, 而离域和杂化泛函则过高地估计了团簇的结构参数. 电子结构分析表明, 中心Au原子与外围的Au原子之间通过 d 电子的成键作用构成团簇内核[Au@Au8]3+, [Au@Au8]3+与PR3配体则通过"σ给予/π反馈"模式成键. PR3配体与[Au@Au8]3+的结合能够加强内核-外围Au原子间的成键作用, 缩小外围Au原子在成键上的差异, 增大前线轨道能级间隙, 从而提高团簇的稳定性. PR3配体中R基团供、 吸电子能力的变化对[Au@Au8(PR3)8]3+结构影响较小, 但对[Au@Au8]3+-PR3结合能影响较大. 能量分析显示, 不同PR3配体与[Au@Au8]3+之间具有相近的轨道作用能, 与R基团供、 吸电子能力相关的非轨道作用能成为影响两者连接牢固程度的决定因素.  相似文献   

2.
仇毅翔  李佳  王曙光 《化学学报》2010,68(7):611-616
采用从头计算MP2方法和密度泛函理论方法, 对过渡金属团簇[PdAu8(PR3)8]2+(R=Me, OMe, H, F, Cl, CN)的几何结构、电子结构以及团簇各组成部分之间的结合能进行了研究. MP2方法和SVWN局域泛函能够对团簇的结构给予准确的描述, 而离域泛函BP86, PBE, BLYP和杂化泛函B3LYP则过高地估计了团簇的几何结构参数. 电子结构研究表明Pd, Au原子通过 d电子的成键作用构成团簇内核[PdAu8]2+, [PdAu8]2+与PR3配体则通过“σ给予/π反馈”模式成键. PR3配体与[PdAu8]2+的结合能够加强Pd-Au之间的成键作用, 增大前线轨道能级间隙, 从而提高团簇的稳定性. PR3配体中R基团供、吸电能力的变化对[PdAu8(PR3)8]2+结构的影响较小, 但对[PdAu8]2+-PR3结合能的影响较大. 能量分析显示不同PR3与[PdAu8]2+之间具有相近的轨道作用能, 与R基团供、吸电能力相关的非轨道作用能成为影响两者连接牢固程度的决定因素.  相似文献   

3.
用ab initio分子轨道方法(RHF,UHF)和密度泛函(DFT)方法研究了团簇Co2S+,Co3S2+的各种可能的几何构型和电子结构,并计算了相应的较稳定构型的振动光谱,发现Co2S+和Co3S2+团簇最稳定结构均具有C,对称性.对团簇的成键作用机理进行了理论分析.  相似文献   

4.
用密度泛函理论(DFT)中的杂化密度泛函B3LYP方法, 在6-31G*水平上对Al8P8团簇的环状结构进行了几何结构优化, 并在同一水平上计算了Al8P8团簇的电子结构、振动特性及极化率和超极化率. 用自然键轨道(NBO)方法分析了成键性质, Al8P8团簇中离子键和共价键共存, 而且在不同轨道中原子间成键有不同的杂化方式. 计算结果表明: 优化后的Al8P8团簇为双层环状结构; 价电子态密度显示其电子结构具有半导体的性质; 最强的IR和Raman谱峰分别位于530.65 cm-1和366. 54 cm-1处.  相似文献   

5.
采用DFT方法研究了在团簇Au20的顶端位点和面心位点配位PH3分子时的几何结构、电子结构以及Au-P的成键机理和能量分析.在两种配位方式下,配体PH3对团簇的几何结构影响都表现为强烈的局域形变效应.不同的配位方式下PH3与团簇的轨道作用方式不同,所形成的团簇化合物电子组态不同.两种配位方式下Au-P成键能的区别主要是来自于配体与团簇之间的Pauli排斥的不同,在面心配位时配体与团簇之间更大的Pauli排斥作用导致了该配位方式的不稳定.  相似文献   

6.
采用从头计算HF,MP2方法和密度泛函理论,对Au(II)系列化合物[Au(CH2)2PH2]2X2(X=F,Cl,Br,I)的几何结构、电子结构和振动频率进行了研究.研究表明Au的5d和6s电子参与Au—Au以及Au—X之间的成键.Au—Au,Au—X键强烈的电子相关作用使HF方法不适于该体系的研究,BP86和B3LYP两种泛函给出较大的Au—Au和Au—X键长,而MP2方法和局域的密度泛函方法则给出了合理的结构参数.局域密度泛函方法计算得到的Au—Au键和Au—X键振动频率也与实验数据符合较好.还运用含时密度泛函理论计算了[Au(CH2)2PH2]2X2的电子激发能,对分子在紫外-可见光谱范围内的电子跃迁进行了分析,考察了卤素配体对激发能的影响,并结合分子轨道能级的变化对此给予了解释.  相似文献   

7.
采用密度泛函理论(DFT)的B3LYP泛函, 在6-311G*水平上对B2Cn+(n=1~9)团簇的几何构型和电子结构进行了优化和振动频率计算. 结果表明, 在B2Cn+(n=1~9)团簇的基态构型中, B2C2+、B2C3+为具有D∞h对称性的线形结构, B2C7+为具有Cs对称性的立体环状结构, 其余均为平面构型; 其成键顺序为C—C成键优于B—C 成键, B—C成键优于B—B成键. 进一步得到了B2Cn+(n=1~9)团簇的总能量(ET)、零点能(EZ)、摩尔热容(Cp)、标准熵(S0)以及原子化能(ΔEn+). 其结果显示, 随着n的递增, ET、EZ、Cp、S0和ΔEn+数值均呈现增大趋势, 其中EZ数值呈现近似等梯度的增加趋势. 通过对B2Cn+(n=1~9)团簇基态结构的垂直电子亲合势的研究发现, n为奇数的B2Cn+团簇比n为偶数的稳定.  相似文献   

8.
仇毅翔  李佳  王曙光 《化学学报》2009,67(14):1585-1590
采用ab initio HF, MP2方法和密度泛函理论方法, 对Pd(0), Pd(I)双核配合物Pd2L2和Pd2L2X2 (L=Me2PCH2PMe2; X=F, Cl, Br, I, H)的几何结构和电子结构进行了研究. 研究表明Pd2L2中Pd原子间的相互作用主要来自电子相关效应, Pd2L2X2中Pd原子间的相互作用则主要来自d轨道的成键作用. MP2方法和局域泛函Xα方法能对两类配合物的几何结构给予准确的描述. 在Pd2L2中, Pd原子的4d电子组成一一对应的成键、反键轨道, 轨道作用相互抵消使Pd原子间仅存在微弱的相互作用. X原子与Pd2L2的作用使Pd—Pd反键轨道电子占据数减少, 成键作用加强. 两类配合物的 Pd—Pd键长与NAO键级之间存在很好的线性关系. 还对Pd2L2和Pd2L2X2的低占据电子激发态进行了含时密度泛函理论计算, 分析不同配合物的电子跃迁特征, 并就卤素配体对Pd2L2X2光谱性质的影响进行了讨论.  相似文献   

9.
用密度泛函理论方法计算了CO分子吸附在有机配体聚乙烯吡咯烷酮poly(N-vinyl-2-pyrrolidone)(PVP)保护下的Au20团簇上的稳定构型的结构和性质。配体PVP通过物理吸附主要作用于Au20团簇的顶点位置。与Au20比较,配体的存在有利于CO的吸附和活化,其根本原因是PVP和CO在Au20表面分别作为供电子和吸电子基团产生的协同效应。中性及阴离子Au20团簇对配体和CO的吸附强度不同,前者对PVP吸附作用较强,后者对CO的吸附和活化作用较强。  相似文献   

10.
用杂化密度泛函B3LYP方法研究了(AB)8(AB=BN,AlP,GaAs,InSb)团簇环形结构的平衡几何构型、电子结构、振动特性以及极化率。计算结果表明,(AB)8团簇的双层环状结构中,每个A(B)原子都与3个B(A)原子成键,且Ⅴ族元素的原子比Ⅲ族元素的原子更接近团簇中心,(BN)8、(AlP)8、(GaAs)8、(InSb)8的平均极化率依次增大,IR和Raman谱峰发生红移。另外,讨论了热力学稳定性和动力学稳定性的变化。  相似文献   

11.
The attempt to prepare hitherto unknown homopolyatomic cations of sulfur by the reaction of elemental sulfur with blue S8(AsF6)2 in liquid SO2/SO2ClF, led to red (in transmitted light) crystals identified crystallographically as S8(AsF6)2. The X-ray structure of this salt was redetermined with improved resolution and corrected for librational motion: monoclinic, space group P2(1)/c (No. 14), Z = 8, a = 14.986(2) A, b = 13.396(2) A, c = 16.351(2) A, beta = 108.12(1) degrees. The gas phase structures of E8(2+) and neutral E8 (E = S, Se) were examined by ab initio methods (B3PW91, MPW1PW91) leading to delta fH theta[S8(2+), g] = 2151 kJ/mol and delta fH theta[Se8(2+), g] = 2071 kJ/mol. The observed solid state structures of S8(2+) and Se8(2+) with the unusually long transannular bonds of 2.8-2.9 A were reproduced computationally for the first time, and the E8(2+) dications were shown to be unstable toward all stoichiometrically possible dissociation products En+ and/or E4(2+) [n = 2-7, exothermic by 21-207 kJ/mol (E = S), 6-151 kJ/mol (E = Se)]. Lattice potential energies of the hexafluoroarsenate salts of the latter cations were estimated showing that S8(AsF6)2 [Se8(AsF6)2] is lattice stabilized in the solid state relative to the corresponding AsF6- salts of the stoichiometrically possible dissociation products by at least 116 [204] kJ/mol. The fluoride ion affinity of AsF5(g) was calculated to be 430.5 +/- 5.5 kJ/mol [average B3PW91 and MPW1PW91 with the 6-311 + G(3df) basis set]. The experimental and calculated FT-Raman spectra of E8(AsF6)2 are in good agreement and show the presence of a cross ring vibration with an experimental (calculated, scaled) stretching frequency of 282 (292) cm-1 for S8(2+) and 130 (133) cm-1 for Se8(2+). An atoms in molecules analysis (AIM) of E8(2+) (E = S, Se) gave eight bond critical points between ring atoms and a ninth transannular (E3-E7) bond critical point, as well as three ring and one cage critical points. The cage bonding was supported by a natural bond orbital (NBO) analysis which showed, in addition to the E8 sigma-bonded framework, weak pi bonding around the ring as well as numerous other weak interactions, the strongest of which is the weak transannular E3-E7 [2.86 A (S8(2+), 2.91 A (Se8(2+)] bond. The positive charge is delocalized over all atoms, decreasing the Coulombic repulsion between positively charged atoms relative to that in the less stable S8-like exo-exo E8(2+) isomer. The overall geometry was accounted for by the Wade-Mingos rules, further supporting the case for cage bonding. The bonding in Te8(2+) is similar, but with a stronger transannular E3-E7 (E = Te) bonding. The bonding in E8(2+) (E = S, Se, Te) can also be understood in terms of a sigma-bonded E8 framework with additional bonding and charge delocalization occurring by a combination of transannular n pi *-n pi * (n = 3, 4, 5), and np2-->n sigma * bonding. The classically bonded S8(2+) (Se8(2+) dication containing a short transannular S(+)-S+ (Se(+)-Se+) bond of 2.20 (2.57) A is 29 (6) kJ/mol higher in energy than the observed structure in which the positive charge is delocalized over all eight chalcogen atoms.  相似文献   

12.
Chemically reasonable models of PR3 (R = Me, Et, iPr, and tBu) were constructed to apply the post Hartree-Fock method to large transition metal complexes. In this model, R is replaced by the H atom including the frontier orbital consistent quantum capping potential (FOC-QCP) which reproduces the frontier orbital energy of PR3. The steric effect is incorporated by the new procedure named steric repulsion correction (SRC). To examine the performance of this FOC-QCP method with the SRC, the activation barriers and reaction energies of the reductive elimination reactions of C2H6 and H2 from M(R1)2(PR2(3))2 (M = Ni, Pd, or Pt; R1 = Me for R2 = Me, Et, or iPr, or R1 = H for R2 = tBu) were evaluated with the DFT[B3PW91], MP4(SDQ), and CCSD(T) methods. The FOC-QCP method reproduced well the DFT[B3PW91]- and MP4(SDQ)-calculated energy changes of the real complexes with PMe3. For more bulky phosphine, the SRC is important to present correct energy change, in which the MP2 method presents reliable steric repulsion correction like the CCSD(T) method because the systems calculated in the SRC do not include a transition metal element. The monomerization energy of [RhCl(PiPr3)2]2 and the coordination energies of CO, H2, N2, and C2H4 with [RhCl(PiPr3)2]2 were theoretically calculated by the CCSD(T) method combined with the FOC-QCP and the SRC. The CCSD(T)-calculated energies agree well with the experimental ones, indicating the excellent performance of the combination of the FOC-QCP with the SRC. On the other hand, the DFT[B3PW91]-calculated energies of the real complexes considerably deviate from the experimental ones.  相似文献   

13.
New organometallic clusters with the MFe2(mu3-S)2 core (M = Mo or Fe) have been synthesized from inorganic [MoFe3S4] or [Fe4S4] clusters under high pressure CO. The reaction of (Cl4-cat)2Mo2Fe6S8(PR3)6[R = Et, (n)Pr] with high pressure CO produced the crystalline [MoFe2S2]4+ clusters, (Cl4-cat)Mo(O)Fe2S2(CO)(n)(PR3)6-n[n= 4, Et =I, (n)Pr =II; n = 5, Et =III] after flash column chromatography. The similar [MoFe2S2]4+ cluster, (Cl4-cat)2MoFe2S2(CO)2(depe)(2)(IV), also has been achieved by the reactions of (Cl4-cat)MoFe3S3(CO)6(PEt3)2 with depe by reductive decoupling of the cluster. For the [Fe3(mu3-S)2]4+ cluster, [Fe4S4(PcHex3)4](BPh4) was reacted with high pressure CO to produce a new Fe3S2(CO)7(PcHex)(2)(V) compound. These reactions generalized the preparation of organometallic compounds from inorganic clusters. All the compounds have been characterized by single crystal X-ray crystallography. A possible reaction pathway for the synthesis of the MFe2(mu3-S) clusters (M = Mo or Fe) has also been suggested.  相似文献   

14.
A comparison of the performance of various density functional methods including long‐range corrected and dispersion corrected methods [MPW1PW91, B3LYP, B3PW91, B97‐D, B1B95, MPWB1K, M06‐2X, SVWN5, ωB97XD, long‐range correction (LC)‐ωPBE, and CAM‐B3LYP using 6‐31+G(d,p) basis set] in the study of CH···π, OH···π, and NH···π interactions were done using weak complexes of neutral (A) and cationic (A+) forms of alanine with benzene by taking the Møller–Plesset (MP2)/6‐31+G(d,p) results as the reference. Further, the binding energies of the neutral alanine–benzene complexes were assessed at coupled cluster (CCSD)/6‐31G(d,p) method. Analysis of the molecular geometries and interaction energies at density functional theory (DFT), MP2, CCSD methods and CCSD(T) single point level reveal that MP2 is the best overall performer for noncovalent interactions giving accuracy close to CCSD method. MPWB1K fared better in interaction energy calculations than other DFT methods. In the case of M06‐2X, SVWN5, and the dispersion corrected B97‐D, the interaction energies are significantly overrated for neutral systems compared to other methods. However, for cationic systems, B97‐D yields structures and interaction energies similar to MP2 and MPWB1K methods. Among the long‐range corrected methods, LC‐ωPBE and CAM‐B3LYP methods show close agreement with MP2 values while ωB97XD energies are notably higher than MP2 values. © 2010 Wiley Periodicals, Inc. J Comput Chem 2010  相似文献   

15.
The CCSD(T) interaction energies for the H‐bonded and stacked structures of the uracil dimer are determined at the aug‐cc‐pVDZ and aug‐cc‐pVTZ levels. On the basis of these calculations we can construct the CCSD(T) interaction energies at the complete basis set (CBS) limit. The most accurate energies, based either on direct extrapolation of the CCSD(T) correlation energies obtained with the aug‐cc‐pVDZ and aug‐cc‐pVTZ basis sets or on the sum of extrapolated MP2 interaction energies (from aug‐cc‐pVTZ and aug‐cc‐pVQZ basis sets) and extrapolated ΔCCSD(T) correction terms [difference between CCSD(T) and MP2 interaction energies] differ only slightly, which demonstrates the reliability and robustness of both techniques. The latter values, which represent new standards for the H‐bonding and stacking structures of the uracil dimer, differ from the previously published data for the S22 set by a small amount. This suggests that interaction energies of the S22 set are generated with chemical accuracy. The most accurate CCSD(T)/CBS interaction energies are compared with interaction energies obtained from various computational procedures, namely the SCS–MP2 (SCS: spin‐component‐scaled), SCS(MI)–MP2 (MI: molecular interaction), MP3, dispersion‐augmented DFT (DFT–D), M06–2X, and DFT–SAPT (SAPT: symmetry‐adapted perturbation theory) methods. Among these techniques, the best results are obtained with the SCS(MI)–MP2 method. Remarkably good binding energies are also obtained with the DFT–SAPT method. Both DFT techniques tested yield similarly good interaction energies. The large magnitude of the stacking energy for the uracil dimer, compared to that of the benzene dimer, is explained by attractive electrostatic interactions present in the stacked uracil dimer. These interactions force both subsystems to approach each other and the dispersion energy benefits from a shorter intersystem separation.  相似文献   

16.
Halogen-hydride interactions between Z-X (Z = CN, NC and X = F, Cl, Br) as halogen donor and H-Mg-Y (Y = H, F, Cl, Br, CH(3)) as electron donor have been investigated through the use of Becke three-parameter hybrid exchange with Lee-Yang-Parr correlation (B3LYP), second-order M?ller-Plesset perturbation theory (MP2), and coupled-cluster single and double excitation (with triple excitations) [CCSD(T)] approaches. Geometry changes during the halogen-hydride interaction are accompanied by a mutual polarization of both partners with some charge transfer occurring from the electron donor subunit. Interaction energies computed at MP2 level vary from -1.23 to -2.99 kJ/mol for Z-F···H-Mg-Y complexes, indicating that the fluorine interactions are relatively very weak but not negligible. Instead, for chlorine- and bromine-containing complexes the interaction energies span from -5.78 to a maximum of -26.42 kJ/mol, which intimate that the interactions are comparable to conventional hydrogen bonding. Moreover, the calculated interaction energy was found to increase in magnitude with increasing positive electrostatic potential on the extension of Z-X bond. Analysis of geometric, vibrational frequency shift and the interaction energies indicates that, depending on the halogen, CN-X···H interactions are about 1.3-2.0 times stronger than NC-X···H interactions in which the halogen bonds to carbon. We also identified a clear dependence of the halogen-hydride bond strength on the electron-donating or -withdrawing effect of the substituent in the H-Mg-Y subunits. Furthermore, the electronic and structural properties of the resulting complexes have been unveiled by means of the atoms in molecules (AIM) and natural bond orbital (NBO) analyses. Finally, several correlative relationships between interaction energies and various properties such as binding distance, frequency shift, molecular electrostatic potential, and intermolecular density at bond critical point have been checked for all studied systems.  相似文献   

17.
The reactions of the previously reported cluster complexes [Re(6)(mu(3)-Se)(8)(PEt(3))(5)I]I, trans-[Re(6)(mu(3)-Se)(8)(PEt(3))(4)I(2)], and cis-[Re(6)(mu(3)-Se)(8)(PEt(3))(4)I(2)] with the [Re(6)(mu(3)-Se)(8)](2+) core with CO in the presence of AgSbF(6) afforded the corresponding cluster carbonyls [Re(6)(mu(3)-Se)(8)(PEt(3))(5)(CO)][SbF(6)](2) (), trans-[Re(6)(mu(3)-Se)(8)(PEt(3))(4)(CO)(2)][SbF(6)](2) (), and cis-[Re(6)(mu(3)-Se)(8)(PEt(3))(4)(CO)(2)][SbF(6)](2) (). Infrared spectroscopy indicated weakening of the bond in CO, suggesting the existence of backbonding between the cluster core and the CO ligand(s). Electrochemical studies focusing on the reversible, one-electron oxidation of the cluster core revealed a large increase in the oxidation potential upon going from the acetonitrile derivatives to their carbonyl analogs, consistent with the depleted electron density of the cluster core upon CO ligation. Disparities between the IR spectra and oxidation potential between and indicate that electronic differences exist between sites trans and cis to the location of a ligand of interest. The active role played by the Se atoms in influencing the cluster-to-CO bonding interactions is suggested through this result and density functional (DF) computational analysis. The computations indicate that molecular orbitals near the HOMO account for backbonding interactions with a high percentage of participation of Se orbitals.  相似文献   

18.
The geometries and interaction energies of stacked and hydrogen-bonded uracil dimers and a stacked adeninecdots, three dots, centeredthymine pair were studied by means of high-level quantum chemical calculations. Specifically, standard as well as counterpoise-corrected optimizations were performed at second-order Moller-Plesset (MP2) and coupled cluster level of theory with single, double, and perturbative triple excitations [CCSD(T)] levels with various basis sets up to the complete basis set limit. The results can be summarized as follows: (i) standard geometry optimization with small basis set (e.g., 6-31G(*)) provides fairly reasonable intermolecular separation; (ii) geometry optimization with extended basis sets at the MP2 level underestimates the intermolecular distances compared to the reference CCSD(T) results, whereas the MP2/cc-pVTZ counterpoise-corrected optimization agrees well with the reference geometries and, therefore, is recommended as a next step for improving MP2/cc-pVTZ geometries; (iii) the stabilization energy of stacked nucleic acids base pairs depends considerably on the method used for geometry optimization, so the use of reliable geometries, such as counterpoise-corrected MP2/cc-pVTZ ones, is recommended; (iv) the density functional theory methods fail completely in locating the energy minima for stacked structures and when the geometries from MP2 calculations are used, the resulting stabilization energies are strongly underestimated; (v) the self-consistent charges-density functional tight binding method, with inclusion of the empirical dispersion energy, accurately reproduces interaction energies and geometries of dispersion-bonded (stacked) complexes; this method can thus be recommended for prescanning the potential energy surfaces of van der Waals complexes.  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