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 156 毫秒
1.
用密度泛函(DFT)方法研究了铜原子簇Cu~n(n=2,3,4,6)的稳定几何构型和电子结构。通过拟合从头算势能面构造铜原子簇势能函数的双体、三体及四体项,并利用该函数和全局优化“Basin-Hopping”算法得到较大铜原子簇(n=13~56)能量极小的结构,计算结果与实验及其它计算结果相一致。  相似文献   

2.
采用密度泛函(DFT)方法结合全局优化"Basin-Hopping"算法研究了铌原子簇: 对于Nb小簇n=2~6我们用密度泛函方法计算了它们的稳定几何构型和电子结构, 通过拟合计算结果构造铌原子簇势能函数, 并利用该函数和全局优化"Basin-Hopping"算法得到较大铌原子簇(n=7~20)能量极小的结构. 计算结果表明与实验及其它计算结果相一致.  相似文献   

3.
本文采用包含Axilord—Teller三体势的分子力学方法,计算了锂原子簇的平衡几何构型,结果表明,锂原子簇的势能面上存在一些近简并的结构。但最稳定结构与从头算的结果基本一致,同时对气相原子簇的生长模式、簇尺寸增大原子簇平均结合能的变化进行了讨论。  相似文献   

4.
本文采用量子化学半经验分子轨道法(CNDO方案),选取了三个晶面层数不同的AgBr晶体模型,对在这些晶体模型(110)表面上的银原子簇Agn(n=1,2,3,4)的电子结构及性质以及潜影机理作了量子化学研究。计算结果表明,晶体对银原子簇的性质有很大的影响,Ag3和Ag4优先取竖立晶面的结构。并且从Ag3起,银原子簇的电荷分布成为上正下负的极性体。本文认为,这种极性是潜影的一个重要性质。Ag3的热稳定性最强,又是银原子簇产生极性的转折点,故Ag3是形成潜影的最小银原子簇。计算表明,银原子簇的增长是离子步骤先于电子步骤。  相似文献   

5.
Ni-P非晶态合金中电子转移问题的DFT研究   总被引:4,自引:1,他引:3  
方志刚  沈百荣  范康年  邓景发 《化学学报》1999,57(11):1246-1251
根据Ni-P非晶态合金结构的短程有序、Ni和P之间具较强化学作用和结构中不存在P-P直接相连的实验事实,选择了Ni~nP(n=1-6)原子簇模型,用密度泛函理论方法对其进行计算。结果表明,在模型体系中,随着P含量的减少,电子转移方向发生变化,P原子由得电子变为失电子。这与Ni-P非晶态合金的实验结果一致,说明Ni~nP(n=1-6)原子簇模型能反映Ni-P非晶态合金的结构特点。  相似文献   

6.
基于从固体锡确立的多体展开势能函数, 采用座标直接优化方法预测了锡原子簇分子(Sn_2-Sn_(300))的结构和相对稳定性, 并用蒙特卡罗方法验证了有关小的原子簇(Sn_2-Sn_(15))的所有结果. 优化结果表明: (1)小的锡原子簇分子(Sn_2-Sn_(15))倾向于密堆积结构, 与锡晶体结构无关; (2)中等大小的簇分子, 如Sn_(47), Sn_(71), Sn_(87)及Sn_(147)等, 则呈α-Sn晶体的畸变结构, 其外围各层原子到中心原子的距离受到压缩, 且某些层被劈裂为两层或多层; (3)随着原子簇尺寸的进一步增大, 结构畸变逐渐减弱, 簇分子的单原子平均结合能缓慢增大, 其外椎值大约在Sn_(740)处趋近于α-Sn的结合能.  相似文献   

7.
根据Co-B非晶态结构的短程有序、Co和B之间是较强的化学作用以及化学键理论,设计了ComB2 (m=1~4) 原子簇模型,用DFT方法对其进行高水平的量子化学计算,结果表明,模型体系ComB2 (m=1~4)中, B原子供给Co原子电子,这与非晶态合金的实验结果一致,同时存在B-B直接相连,为了比较,也选择了ConB (n=1~4)模型,计算结果与实验不符,说明ComB2(m=1~4)原子簇模型更能反映非晶态的结构特点.  相似文献   

8.
本文报道六核锇排状平面原子簇的合成和活化。三核锇活性原子簇[Os_3(CO)_(10)(MeCN)_2]在PdCl_2催化作用下发生偶联反应,生成具有活性的六核锇排状原子簇[Os_6(CO)_(20)(MeCN)]和[Os_6(CO)_(19)(MeCN)_2]。用氧化三甲胺(Me_3NO)对活性原子簇中羰基配位基进行多重活化,获得一系列具有活性的六核锇排状原子簇[Os_6(CO)_(21-n)(MeCN)_n](n=1—5)及其相应的配位体取代反应产物[Os_6(CO)_(21-n)(P(OMe)_3)_n](n=1—5)。并通过红外光谱、核磁共振波谱和质谱等数据对上述排状原子簇进行结构表征,同时就其排状平面结构和反应性能特点进行讨论。  相似文献   

9.
Ni-B非晶态合金局域结构和电荷转移性质的理论研究   总被引:3,自引:1,他引:2  
根据非晶态合金结构的短程有序、Ni和B间存在较强的化学作用、结构中存在B-B键直接相连的实验事实,选择Ni~mB~2(m=1,2,4)原子簇作为非晶态局域结构的计算模型。考虑原子簇间的相互作用,又对[Ni~4B~2]~n(n=1,2,5,7)簇团进行了系统计算。结果表明在所选Ni-Br的簇模型中,都是B原子提供电荷给Ni原子,这些均与非晶态合金实验结果和一些理论计算结果相符。簇团的计算结果还表明,小原子簇Ni~4B~2内原子间存在较强的化学键作用,而簇间的相互作用相对较弱,很容易造成对称性破缺而导致产生长程无序,因此Ni-B非晶态可以被看成是由大量i上原子簇无序堆砌成的,这一点也同实验事实相吻合。计算结果也表明我们前面工作中所取的和最常见的Ni-B非晶态合金Ni~6~4B~3~6有相似组成的Ni~4B~2原子簇模型能在一定程度上反映Ni-B非晶态合金的局域结构特征。  相似文献   

10.
多体展开势能函数研究表明,Si4-Si16原子簇分子间的结构衍生关系为:依次增加一个二配位或三配位的表面原子,分子表面被四元蝶形环Si4(D2d)所覆盖;Sin(n=5-16)结构中多含有Si5(D3h)、Si6(D2d)区域结构单元,笼状Si10及Si16的表面原子均为三配位或三配位以上,预计Si5、Si6、Si10及Si16是硅原子簇碎片化产物分布中丰度较高的序列;在这一范围内的分子结构呈与晶体硅结构(金刚石)无关的密堆积,最高配位数为5,在小于半球的立体角内形成六配位的硅中心,使该簇合物在能量上不稳定。  相似文献   

11.
Earlier studies have shown that the most stable structures for (ZnS)n clusters with n = 10-47 are hollow polyhedral clusters ("bubbles"). We report a detailed study of larger clusters, where n = 50, 60, 70, and 80, for which onionlike or "double bubble" structures are predicted. We report calculations of the vibrational spectra and the electronic structure of bubble and double bubble clusters, which may assist in their experimental identification.  相似文献   

12.
The stability and structures of titanium-doped gold clusters Au(n)Ti (n=2-16) are studied by the relativistic all-electron density-functional calculations. The most stable structures for Au(n)Ti clusters with n=2-7 are found to be planar. A structural transition of Au(n)Ti clusters from two-dimensional to three-dimensional geometry occurs at n=8, while the Au(n)Ti (n=12-16) prefer a gold cage structure with Ti atom locating at the center. Binding energy and second-order energy differences indicate that the Au(14)Ti has a significantly higher stability than its neighbors. A high ionization potential, low electron affinity, and large energy gap being the typical characters of a magic cluster are found for the Au(14)Ti. For cluster-cluster interaction between magic transition-metal-doped gold clusters, calculations were performed for cluster dimers, in which the clusters have an icosahedral or nonicosahedral structure. It is concluded that both electronic shell effect and relative orientation of clusters are responsible for the cluster-cluster interaction.  相似文献   

13.
14.
Simulated annealing Monte Carlo conformer searches using the "mag-walking" algorithm are employed to locate the global minima of molecular clusters of ammonium chloride of the types (NH(4)Cl)(n), (NH(4)(+))(NH(4)Cl)(n), and (Cl(-))(NH(4)Cl)(n) with n = 1-13. The M06-2X density functional theory method is used to refine and predict the structures, energies, and thermodynamic properties of the neutral, cation, and anion clusters. For selected small clusters, the resulting structures are compared to those obtained from a variety of models and basis sets, including RI-MP2 and B3LYP calculations. M06-2X calculations predict enhanced stability of the (NH(4)(+))(NH(4)Cl)(n) clusters when n = 3, 6, 8, and 13. This prediction corresponds favorably to anomalies previously observed in thermospray mass spectroscopy experiments. The (NH(4)Cl)(n) clusters show alternations in stability between even and odd values of n. Clusters of the type (Cl(-))(NH(4)Cl)(n) display a magic number distribution different from that of the cation clusters, with enhanced stability predicted for n = 2, 6, and 11. None of the observed cluster structures resemble the room-temperature CsCl structure of NH(4)Cl(s), which is consistent with previous work. Numerous clusters have structures reminiscent of the higher-temperature, rock-salt phase of the solid ammonium halides.  相似文献   

15.
Pyridine containing water clusters, H(+)(pyridine)(m)(H(2)O)(n), have been studied both experimentally by a quadrupole time-of-flight mass spectrometer and by quantum chemical calculations. In the experiments, H(+)(pyridine)(m)(H(2)O)(n) with m = 1-4 and n = 0-80 are observed. For the cluster distributions observed, there are no magic numbers, neither in the abundance spectra, nor in the evaporation spectra from size selected clusters. Experiments with size-selected clusters H(+)(pyridine)(m)(H(2)O)(n), with m = 0-3, reacting with D(2)O at a center-of-mass energy of 0.1 eV were also performed. The cross-sections for H/D isotope exchange depend mainly on the number of water molecules in the cluster and not on the number of pyridine molecules. Clusters having only one pyridine molecule undergo D(2)O/H(2)O ligand exchange, while H(+)(pyridine)(m)(H(2)O)(n), with m = 2, 3, exhibit significant H/D scrambling. These results are rationalized by quantum chemical calculations (B3LYP and MP2) for H(+)(pyridine)(1)(H(2)O)(n) and H(+)(pyridine)(2)(H(2)O)(n), with n = 1-6. In clusters containing one pyridine, the water molecules form an interconnected network of hydrogen bonds associated with the pyridinium ion via a single hydrogen bond. For clusters containing two pyridines, the two pyridine molecules are completely separated by the water molecules, with each pyridine being positioned diametrically opposite within the cluster. In agreement with experimental observations, these calculations suggest a "see-saw mechanism" for pendular proton transfer between the two pyridines in H(+)(pyridine)(2)(H(2)O)(n) clusters.  相似文献   

16.
Bisulfate water clusters, HSO(4)(-)(H(2)O)(n), have been studied both experimentally by a quadrupole time-of-flight mass spectrometer and by quantum chemical calculations. For the cluster distributions studied, there are some possible "magic number" peaks, although the increase in abundance compared to their neighbours is small. Experiments with size-selected clusters with n = 0-25, reacting with D(2)O at a center-of-mass energy of 0.1 eV, were performed, and it was observed that the rate of hydrogen/deuterium exchange is lower for the smallest clusters (n < 8) than for the larger (n > 11), with a transition taking place in the range n = 8-11. We propose that the protonic defect of the bisulfate ion remains rather stationary unless the degree of hydration reaches a given level. In addition, it was observed that H/D scrambling becomes close to statistically randomized for the larger clusters. Insight into this size dependency was obtained by B3LYP/6-311++G(2d,2p) calculations for HSO(4)(-)(H(2)O)(n) with n = 0-10. In agreement with experimental observations, these calculations suggest pronounced effectiveness of a 'see-saw mechanism' for pendular proton transfer with increasing HSO(4)(-)(H(2)O)(n) cluster size.  相似文献   

17.
DFT/TDDFT calculations have been carried out for a series of silver and gold nanorod clusters (Ag(n), Au(n), n = 12-120) whose structures are of cigar-type. Pentagonal Ag(n) clusters with n = 49-121 and hexagonal Au(n) clusters with n = 14-74 were also calculated for comparison. Metal-metal distances, binding energies per atom, ionization potentials, and electron affinities were determined, and their trends with cluster size were examined. The TDDFT calculated excitation energies and oscillator strengths were fit by a Lorentz line shape modification, which gives rise to the simulated absorption spectra. The significant features of the experimental spectra for actual silver and gold nanorod particles are well reproduced by the calculations on the clusters. The calculated spectral patterns are also in agreement with previous theoretical results on different-type Ag(n) clusters. Many differences in the calculated properties are found between the Ag(n) and Au(n) clusters, which can be explained by relativistic effects.  相似文献   

18.
We employ recent flexible ab initio potential energy and dipole surfaces [Y. Wang, X. Huang, B. C. Shepler, B. J. Braams, and J. M. Bowman, J. Chem. Phys. 134, 094509 (2011)] to the calculation of IR spectra of the intramolecular modes of water clusters. We use a quantum approach that begins with a partitioned normal-mode analysis of perturbed monomers, and then obtains solutions of the corresponding Schro?dinger equations for the fully coupled intramolecular modes of each perturbed monomer. For water clusters, these modes are the two stretches and the bend. This approach is tested against benchmark calculations for the water dimer and trimer and then applied to the water clusters (H(2)O)(n) for n = 6-10 and n = 20. Comparisons of the spectra are made with previous ab initio harmonic and empirical potential calculations and available experiments.  相似文献   

19.
Reactions of neutral V(n), Nb(n), and Ta(n) metal clusters (n< or =11) with CO+H(2) mixed gases and CH(3)OH in a flow tube reactor (1-50 Torr) are studied by time of flight mass spectroscopy and density functional theory calculations. Metal clusters are generated by laser ablation, and reactants and products are ionized by low fluence (approximately 200 microJ/cm(2)) 193 nm excimer laser light. Nb(n) clusters exhibit strong size dependent reactivity in reactions both with CO+H(2) and CH(3)OH compared with V(n) and Ta(n) clusters. A "magic number" (relatively intense) mass peak at Nb(8)COH(4) is observed in the reaction of Nb(n) clusters with CO+H(2), and CH(3)OH is suggested to be formed. This feature at Nb(8)COH(4) remains the most intense peak independent of the relative concentrations of CO and H(2) in the flow tube reactor. No other Nb(n), Ta(n), or V(n) feature behaves in this manner. In reactions of CH(3)OH with metal clusters M(n) (M=V, Nb, and Ta, n=3-11), nondehydrogenated products M(n)COH(4)/M(n)CH(3)OH are only observed on Nb(8) and Nb(10), whereas dehydrogenated products M(n)CO/CM(n)O are observed for all other clusters. These observations support the suggestion that CH(3)OH can be formed on Nb(8) in the reaction of Nb(n) with CO+H(2). A reaction mechanism is suggested based on the experimental results and theoretical calculations of this work and of those in the literature. Methanol formation from CO+H(2) on Nb(8) is overall barrierless and thermodynamically and kinetically favorable.  相似文献   

20.
Density functional theory geometry optimizations and reduction potential calculations are reported for all five known oxidation states of [Fe(4)S(4)(SCH(3))(4)](n)()(-) (n = 0, 1, 2, 3, 4) clusters that form the active sites of iron-sulfur proteins. The geometry-optimized structures tend to be slightly expanded relative to experiment, with the best comparison found in the [Fe(4)S(4)(SCH(3))(4)](2)(-) model cluster, having bond lengths 0.03 A longer on average than experimentally observed. Environmental effects are modeled with a continuum dielectric, allowing the solvent contribution to the reduction potential to be calculated. The calculated protein plus solvent effects on the reduction potentials of seven proteins (including high potential iron proteins, ferredoxins, the iron protein of nitrogenase, and the "X", "A", and "B" centers of photosystem I) are also examined. A good correlation between predicted and measured absolute reduction potentials for each oxidation state of the cluster is found, both for relative potentials within a given oxidation state and for the absolute potentials for all known couples. These calculations suggest that the number of amide dipole and hydrogen bonding interactions with the Fe(4)S(4) clusters play a key role in modulating the accessible redox couple. For the [Fe(4)S(4)](0) (all-ferrous) system, the experimentally observed S = 4 state is calculated to lie lowest in energy, and the predicted geometry and electronic properties for this state correlate well with the EXAFS and M?ssbauer data. Cluster geometries are also predicted for the [Fe(4)S(4)](4+) (all-ferric) system, and the calculated reduction potential for the [Fe(4)S(4)(SCH(3))(4)](1)(-)(/0) redox couple is in good agreement with that estimated for experimental model clusters containing alkylthiolate ligands.  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