A combined stochastic search and density functional theory study on the neutral and charged silicon-based clusters MSi6 (M = La, Ce, Yb and Lu)

Huai-Qian Wang* and Hui-Fang Li
College of Engineering, Huaqiao University, Quanzhou, 362021, China. E-mail: hqwang@hqu.edu.cn

Received 25th April 2014 , Accepted 11th June 2014

First published on 11th June 2014


Abstract

The neutral and charged silicon clusters Si6q (q = 0, −1) are doped systematically with an external atom of different lanthanides: La, Ce, Yb and Lu. The structural, electronic, and magnetic properties of the doped clusters, MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) are investigated using the Saunders “Kick” global stochastic search technique combined with density functional theory (DFT) calculations. The accuracy and consistency of structure optimization of MSi6q clusters are tested with three exchange correlation functionals (PBEPBE, BPW91 and B3LYP). DFT calculations show that the pentagonal bipyramid structures of MSi6 (M = Yb, Lu) and Si6/7 with high symmetry are ground state structures, which give simulated photoelectron spectra in good agreement with available experiments. The detailed comparison between previously published experimental photoelectron spectra, and the present theoretical simulation helps to identify the ground state structures. The neutral species shows the impurity as a four-coordinate atom in the equatorial plane of pentagonal bipyramid. Furthermore, the current magnetic property analysis indicates that the magnetic moments of the doped rare earth atoms remain largely localized, and the atomic-like magnetism is maintained in the doped clusters.


1. Introduction

Since silicon is the keystone of modern semiconductor technology, it attracts great attention as building blocks for silicon-based nanomaterials. Silicon clusters have been the subject of extensive experimental1–6 and theoretical7–12 studies in the past two decades. It was observed that these clusters tend to form compact, diamond-like structures rather than the fullerene-like cages that are characteristic of their congener, carbon.13 These results are ascribed to the fact that silicon atoms generally prefer sp3 hybridization to sp2. By the doping of a suitable metal atom into a pure Si cluster, a highly stabilized cluster will open up a new avenue to cluster-assembled materials made from Si atoms. More significantly, however, the lion's share of that interest stems from the realization that the addition of “impurity” atoms can significantly alter the properties of silicon. For example, the notion of altering the electronic properties of silicon by doping group IIIA (Ga, Al) and group VA (P, As) elements into it has been in use for over half a century and has led to the revolutionary rise of modern computers. As is well known, the first hints that silicon cages may be stabilized by the introduction of a doping atom emerged from the early, pioneering experimental work of Beck who observed enhanced ion intensities of metal–silicon clusters, MSin+ (M = Mo, Cr, W) with around n = 16, when a transition metal atom was added to the cluster.14 Since then, many stable MSin species have been reported, in which various factors accounting for their formation have been examined experimentally15–23 and theoretically.24–34 Recently, infrared multiple photon dissociation (IR-MPD) and mass spectrometry have been used to identify the structures of small neutral vanadium- and manganese-doped silicon clusters SinX (n = 6–9, X = V and Mn).15 Recent theoretical work on doping silicon anion clusters with yttrium suggests that the following two growth behaviors play an important role: (1) the metal atom acting as a linker between two subclusters and (2) the metal atom being endohedrally encapsulated in a silicon cage.25

The rare earths (RE) are special transition metals, possessing many of the same properties (such as optical property, magnetism) as the transition elements. RE compounds are usually used as catalysts and synthetic products in the production of petroleum. Up to now, only a few reports exist on RE-doped silicon clusters.17,25,29,35–37 However, interest in their potential applications has spurred considerable activity over the past couple of years. We have recently published the results of comparing experimental photoelectron spectroscopy and theoretical density functional theory (DFT) studies of MSi6−/0 (M = Pr, Gd, Ho) clusters.35 This article provides an interesting example for evaluating the accuracies of various DFT methods, in general, used to verify the proposed structural types for these species. Experimentally, the validity of the concept can be proven by using photoelectron spectroscopy, which is shown to be a powerful experimental technique for probing electronic structures of size-selected clusters.38–45 Especially, Bowen and co-workers have investigated the lanthanide–silicon cluster anions LnSin (3 ≤ n ≤ 13; Ln = Ho, Gd, Pr, Sm, Eu, Yb)46 and EuSin (3 ≤ n ≤ 17)17 by using photoelectron spectroscopy. They proposed that the building blocks of silicon-based clusters may exist as magnetic materials. They also proposed that the Eu atom encapsulated into the Sin clusters at the size n = 12, which has been confirmed by theoretical calculation.29 In addition, Nakajima and co-workers have previously reported the electronic properties of silicon clusters, containing a transition or lanthanide metal atom from group 3, 4, or 5, by using anion photoelectron spectroscopy at 213 nm.47

Here, we report the generation of a series of anionic and neutral lanthanide metal-doped silicon clusters MSi6 (M = La, Ce, Yb and Lu) using Saunders “Kick” global stochastic search technique combined with DFT geometric optimization. The aims of the present work are as follows: (1) to obtain the various global minimum structures of MSi6q clusters; and (2) to compare the results of our extensive computations with previously experimental findings.46,47 Good agreement between previously measured photoelectron spectroscopy and theoretical simulations helps to identify the ground states of the anionic and neutral RE-doped MSi6 clusters. The rest of this article is organized as follows. We describe our computational details and theoretical backgrounds in Section 2. In Section 3, we introduce and compare various geometric structures, simulated photoelectron spectra (PES), as well as electronic and magnetic properties. Finally, the conclusions are given in Section 4.

2. Computational methods

The calculations in this work began with unbiased searches for the low-lying structures of a series of anionic and neutral RE-doped silicon clusters MSi6 (M = La, Ce, Yb and Lu) by using the Saunders “Kick” (SK) global search technique48 combined with DFT geometric optimization (SK-DFT) within the generalized gradient approximation in the Perdue–Burke–Ernzerhof (PBE)49 functional form using the GAUSSIAN09 program.50 This SK stochastic method generates structures randomly and facilitates the thorough exploration of unknown isomers much more easily than manual methods. Specifically, all the atoms are placed at the same point initially, and they are then “kicked” randomly within a sphere of radius R (R is the maximum kick distance), e.g., a hollow sphere of inside radius 10 Å. The method developed further in this work generates up to 500 unbiased starting geometries, and then submits them automatically for optimization with GAUSSIAN09. We have recently successfully employed the SK-DFT method for global minimum searches on a series of bimetallic cluster and vanadium oxides.51–57 The scheme of this research consists of the following four steps. Firstly, the initial SK global minimum search was performed using the DFT method known in the literature as “PBEPBE” functional49 with a scalar relativistic effective core potential and a LANL2DZ basis set for the RE atom and a 6–31G basis set for Si atom (the PBEPBE/RE/LANL2DZ/Si/6–31G level), which is implemented in the GAUSSIAN09 package. Many structural isomers were generated for each cluster. The kick method runs at the PBEPBE/RE/LANL2DZ/Si/6–31G level some 500 times until no new minima appeared. If this procedure is repeated enough number of times, eventually all isomeric structures for the cluster will be found. Secondly, the distinct isomers were ranked according to their relative energies at the PBEPBE/RE/LANL2DZ/Si/6–31G level. Moreover, the effect of spin multiplicity is also taken into account in the present study. Thirdly, we selected the top-eight lowest-energy isomers for each species, and those with their energy value within ∼1.50 eV from the lowest-energy isomer are to be further optimized using the PBEPBE exchange–correlation functional with the larger RE/SDD/Si/6–311+G(d) basis sets (the PBEPBE/RE/SDD/Si/6–311+G(d) level). Here, the SDD basis set58 along with the quasi-relativistic MWB (28 core electrons) pseudopotential58,59 was chosen for RE atoms. For transition metals compounds, lanthanides and actinides, the SDD basis set is standard practice, it is slightly superior and usually gives very nice structures.60 For Si, we adopted the 6–311+G(d) basis set. Such diffuse functions improve the predicted properties of species with extended electronic densities such as anions. Finally, the top-several isomers were ranked again by their relative energies at high PBEPBE/RE/SDD/Si/6–311+G(d) level, and those with their energy value within 0.40 eV from the lowest-energy isomer were all regarded as potential candidate lowest-energy structures. All the optimization procedure was carried out without any symmetry constraints. This was followed by frequency calculations to ensure that all found structures are at local minima of the potential energy surface and have real frequencies.

The reliability of the present computational method (at PBEPBE/RE/SDD/Si/6–311+G(d) level) was validated by performing calculations on the features of simulated PES spectra and the first adiabatic/vertical detachment energy (ADE/VDE) for which experimental results are available (see Fig. 2, 5, 6 and Table 3).4,6,46,47 Furthermore, geometry optimization of MSi6 (M = La, Ce, Yb and Lu) clusters were tested with several exchange–correlation functionals (PBEPBE,49 BPW9161,62 and B3LYP63–65) for accuracy and consistency. The calculated results using these three different DFT levels are presented in Table 2. These results indicate that the geometries of all the clusters obtained by three functionals are similar, although the order of isomers is reversed in some cases. Here, only PBEPBE results will be discussed in the text. In addition, due to the possible multireference structure of wavefunction, the DFT functional we employed here may be insufficient in giving a very accurate description. The multi-reference configuration interaction (MRCI) method, which is based on multiconfiguration self-consistent field (MC-SCF) wavefunction, would be more optimal for the present clusters. Unfortunately, it is currently not computationally feasible to study the RE-doped clusters using this advanced theoretical method. Geometries are regarded as optimized when the maximum force, the root-mean-square (RMS) force, the maximum displacement of atoms, and the RMS displacement of atoms have magnitudes less than 0.00045, 0.0003, 0.0018, and 0.0012 a.u., respectively. Because the computed zero-point energy (ZPE) corrections of the isomers of a specific cluster size are small and almost the same, they are not expected to affect the relative energy ordering. Hence, the relative energies of isomers are obtained from the total electronic energies (excluding ZPE).

We used three criteria in comparing the theoretical results with the experimental data to select our most likely candidate structures: (1) the relative energies; (2) the first VDE and ADE; (3) the number of distinct peaks of simulated PES in the low binding energy range of ≤266 nm (4.661 eV) and their relative positions. The first criterion addresses typical errors in DFT calculations, which is about several tenths of an electron volt in relative energies for MSi6q clusters. In the present study, we used a cut-off energy of 0.40 eV at the PBEPBE/RE/SDD/Si/6–311+ G(d) level to obtain low-lying isomers. Next, we calculated the first VDE (which is defined as the energy difference between the neutral clusters at optimized anion geometry clusters and optimized anion clusters: VDE = Eneutral at optimized anion geometryEoptimized anion) and ADE (which is the energy difference between the optimized anion geometry and the optimized neutral geometry: ADE = Eoptimized neutralEoptimized anion). Finally, the simulated spectra were then compared to the measured PES spectra to distinguish from top candidates for the low-lying isomers. VDEs were calculated using the generalized Koopmans' theorem by adding a correction term to the eigenvalues of the anion.66 The correction term was calculated as δE = E1E2εHOMO, where E1 and E2 are the total energies of the anion and neutral, respectively, in their ground states at the anion equilibrium geometry and εHOMO corresponds to the eigenvalue of the highest occupied molecular orbital (HOMO) of the anion. The simulated PES spectra were fitted with a full width at half-maximum (FWHM) of 0.20 eV.53 These methods have been used successfully in a number of previous studies, and they have been shown to yield VDEs in good agreement with PES data.38–45 We paid particular attention to the number of peaks below 4.661 eV. Only those isomers that meet all three criteria are considered to be the superlative candidates for MSi6 clusters.

3. Results and discussions

The optimized anion and neutral ground-state and low-lying structures at the PBEPBE/RE/SDD/Si/6–311+G(d) level are presented in Fig. 1 for both pure Si6/7q and RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters. Table 1 gives various structural and energetic characteristics of parent Si6/7q and doped MSi6q clusters. The calculated relative energies for the top-eight lowest-energy isomers of MSi6 (M = La, Ce, Yb and Lu) with three density functional methods are presented in Table 2. The calculated first ADE and VDE of the anionic top-several isomers, as well as some experiment results, are given in Table 3. Natural charges populations of the lowest-energy RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters are listed in Table 4. Table 5 gives the magnetic moment (μB) of 4f, 5d, 6s, and 6p states for RE atom, total magnetic moment (μB) of the RE atom, and total magnetic moment of the lowest-energy RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters. Fig. 2–6 show the simulated PES spectra of the top-several isomers (within ∼0.40 eV), along with experimental PES spectra for comparison. The calculated ADEs are obtained by optimizing the geometry of the neutrals with the ground state geometries of the anions as starting points. The HOMO (SOMO for open-shell species) and the Mülliken spin density pictures of RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters are presented in Fig. 7–10.
image file: c4ra03788j-f1.tif
Fig. 1 Equilibrium geometries of the lowest-energy isomers for pure Si6q (Si7q) (q = 0, −1) clusters and RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters at the PBEPBE/RE/SDD/Si/6–311+G(d) level of theory. The dark cyan spheres stand for silicon atoms and the black ones for RE atoms.
Table 1 Various structures with the symmetry type (Sym), the spin multiplicity (SM), the binding energy (BE) per atom, the doping energy (DE) per atom, HOMO–LUMO energy gap Egap, and energy separation of the isomeric structures from the lowest-energy structures (ΔE) of both pure Si6/7q (q = 0, −1) clusters and MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters
Cluster The lowest-energy isomer Low lying isomer
Isomer SM Sym BE (eV) DE (eV) Egap (eV) Isomer ΔE (eV) Isomer ΔE (eV)
Si6 c 1 C2v 3.42   2.15 d 0.62 e 1.45
Si6 a 2 C2v 3.55   0.55 b 0.19 d 0.85
Si7 f 1 D5h 3.56 4.37 2.12 g 0.96 h 0.97
Si7 f 2 D5h 3.62 4.04 0.59 g 0.79 h 0.87
LaSi6 i 2 C1 3.59 4.59 0.32 m 0.29 k 0.30
LaSi6 j 1 C5v 3.82 4.45 1.04 k 0.29 l 0.34
CeSi6 i 3 Cs 5.35 16.93 0.17 l 0.39 k 0.42
CeSi6 j 2 Cs 5.63 16.73 0.49 k 0.34 l 0.37
YbSi6 i 1 C2v 3.16 1.58 0.62 o 0.71 p 0.84
YbSi6 i 2 C2v 3.67 1.17 0.42 j 0.34 p 0.56
LuSi6 i 2 C2v 3.35 2.90 0.68 m 0.47 j 0.51
LuSi6 j 1 C5v 3.81 3.01 1.36 i 0.35 l 0.53

Cluster Low lying isomer
Isomer ΔE (eV) Isomer ΔE (eV) Isomer ΔE (eV) Isomer ΔE (eV) Isomer ΔE (eV)
LaSi6 l 0.34 o 0.61 p 0.62 j 0.63 n 0.75
LaSi6 m 0.36 n 0.65 o 0.72 i 0.82 p 1.44
CeSi6 m 0.48 j 0.62 o 0.75 n 0.86 p 0.91
CeSi6 m 0.38 i 0.62 n 0.71 o 0.86 p 1.35
YbSi6 j 0.90 l 0.92 n 0.95 m 0.96 k 1.19
YbSi6 o 0.63 l 0.75 m 0.76 k 0.83 n 1.02
LuSi6 l 0.52 k 0.60 m 0.64 n 0.93 o 0.95
LuSi6 m 0.53 k 0.63 o 0.79 p 0.95 n 1.02


Table 2 Relative energies in eV for selected low-lying isomers of MSi6 (M = La, Ce, Yb and Lu) with different density functional methods
Cluster Isomer Sym PBE B3LYP BPW91 Cluster Isomer Sym PBE B3LYP BPW91
LaSi6 i C1 0.00 0.00 0.00 YbSi6 i C2v 0.00 0.00 0.00
j Cs 0.63 0.63 0.62 j Cs 0.90 0.77 0.88
k C1 0.30 0.28 0.28 k C1 1.19 1.15 1.18
l C1 0.34 0.32 0.34 l C1 0.92 0.55 0.92
m Cs 0.29 0.24 0.27 m Cs 0.96 0.74 0.93
n Cs 0.75 0.65 0.72 n Cs 0.95 0.74 0.93
o C3v 0.61 0.51 0.56 o C3 0.71 0.48 0.65
p Cs 0.62 0.79 0.81 p C2v 0.84 0.66 0.82
CeSi6 i Cs 0.00 0.00 0.00 LuSi6 i C2v 0.00 0.00 0.00
j C1 0.62 0.41 0.69 j Cs 0.51 0.50 0.51
k C1 0.42 0.09 0.42 k C1 0.60 0.61 0.60
l Cs 0.39 0.22 0.40 l Cs 0.52 0.47 0.51
m Cs 0.48 0.15 0.39 m Cs 0.47 0.48 0.47
n Cs 0.86 0.45 0.85 n Cs 0.92 0.84 0.90
o C1 0.75 0.34 0.70 o Cs 0.95 0.88 1.00
p Cs 0.91 0.70 0.87 p Cs 0.64 0.52 0.83


Table 3 Experimentally measured adiabatic and vertical detachment energies (ADEs and VDEs) from the photoelectron spectra compared to the calculated ground state or low-lying anion state ADEs and VDEs for pure Si6/7 and RE-doped MSi6 (M = La, Ce, Yb and Lu) clusters. All energies are in eV
Cluster Isomer RE ADEa,b VDEa
a Numbers in parentheses represent the experimental uncertainties in the last digit.b Electron affinity of the neutral species.c From ref. 4.d From ref. 6.e From ref. 46.f From ref. 47.
Si6 a 0.00 2.24 2.45 (2.39 ± 0.01)c
b 0.19 2.05 2.77
Si7 f 0.00 1.91 (1.85 ± 0.02)d 2.30 (2.39 ± 0.04)c
LaSi6 j 0.00 2.73 2.81
k 0.29 2.11 2.18
l 0.34 2.10 2.19
m 0.36 2.03 2.08
CeSi6 j 0.00 2.66 2.83
k 0.34 2.12 2.19
l 0.37 2.06 2.16
m 0.38 2.09 2.13
YbSi6 i 0.00 1.83 (1.80 ± 0.05)e 1.95 (1.98)e
j 0.34 2.39 2.42
LuSi6 j 0.00 2.86 (2.10)f 3.07 (3.20)f
i 0.35 2.00 2.16


Table 4 Population of the natural charges of the lowest-energy RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters
Cluster Isomer RE Si-1 Si-2 Si-3 Si-4 Si-5 Si-6
a Atoms bonding to RE atom.
LaSi6 i 0.47 −0.10a −0.21a −0.10a −0.21a 0.07 0.08
LaSi6 j 0.16 −0.17a −0.17a −0.17a −0.17a −0.17a −0.29
CeSi6 i 0.61 −0.12a −0.28a −0.12a −0.26a 0.09 0.09
CeSi6 j 0.10 −0.17a −0.17a −0.17a −0.16a −0.16a −0.28
YbSi6 i 0.91 −0.21a −0.34a −0.21a −0.34a 0.09 0.09
YbSi6 i 0.33 −0.26a −0.36a −0.26a −0.36a −0.05 −0.05
LuSi6 i 0.56 −0.10a −0.30a −0.10a −0.30a 0.13 0.13
LuSi6 j 0.30 −0.19a −0.19a −0.19a −0.19a −0.19a −0.33


Table 5 Magnetic moment (μB) of 4f, 5d, 6s, and 6p states for RE atom, total magnetic moment (μB) of the RE atom, and total magnetic moment of the lowest-energy RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters
Cluster Isomer RE Moment (μB) Molecule (μB)
4f 5d 6s 6p Total
LaSi6 i 0.02 0.51 0.09 0.01 0.63 1
LaSi6 j 0.00 0.00 0.00 0.00 0.00 0
CeSi6 i 0.87 0.31 0.52 0.07 1.77 2
CeSi6 j 1.02 0.04 0.00 0.00 1.06 1
YbSi6 i 0.00 0.00 0.00 0.00 0.00 0
YbSi6 i 0.01 0.01 0.67 0.17 0.86 1
LuSi6 i 0.00 0.10 0.59 0.16 0.85 1
LuSi6 j 0.00 0.00 0.00 0.00 0.00 0



image file: c4ra03788j-f2.tif
Fig. 2 Comparison of the experimental photoelectron spectra of Si6 and Si7 clusters with their simulated spectra. Each VDE was fitted with a full width at half-maximum (FWHM) of 0.20 eV to obtain the simulated PES spectra. The experimental PES spectra are cited from ref. 4.

image file: c4ra03788j-f3.tif
Fig. 3 Structures, relative energies, and simulated photoelectron spectra for the four low-lying isomers of LaSi6. Each VDE was fitted with a full width at half-maximum (FWHM) of 0.20 eV to obtain the simulated PES spectra.

image file: c4ra03788j-f4.tif
Fig. 4 Structures, relative energies, and simulated photoelectron spectra for the four low-lying isomers of CeSi6. Each VDE was fitted with a full width at half-maximum (FWHM) of 0.20 eV to obtain the simulated PES spectra.

image file: c4ra03788j-f5.tif
Fig. 5 Structures, relative energies, and simulated photoelectron spectra for the two low-energy isomers of YbSi6. Each VDE was fitted with a full width at half-maximum (FWHM) of 0.20 eV to obtain the simulated PES spectra. The black curve is for the experimental PES for YbSi6. The experimental PES spectra are cited from ref. 46.

image file: c4ra03788j-f6.tif
Fig. 6 Structures, relative energies, and simulated photoelectron spectra for the two low-energy isomers of LuSi6. Each VDE was fitted with a full width at half-maximum (FWHM) of 0.20 eV to obtain the simulated PES spectra. The black curve is for the experimental PES for LuSi6. The experimental PES spectra are cited from ref. 47.

image file: c4ra03788j-f7.tif
Fig. 7 (a) The HOMO picture of LaSi6 (C5v, 1A′). (b) The SOMO picture of LaSi6 (C1, 2A). (c) The Mülliken spin density of LaSi6 (C1, 2A).

image file: c4ra03788j-f8.tif
Fig. 8 (a) The SOMO picture of CeSi6 (Cs, 2A′′). (b) The SOMO and (c) SOMO-1 pictures of CeSi6 (Cs, 3A′′). (d) The Mülliken spin density of CeSi6 (Cs, 2A′′). (e) The Mülliken spin density of CeSi6 (Cs, 3A′′).

image file: c4ra03788j-f9.tif
Fig. 9 (a) The SOMO picture of YbSi6 (C2v, 2A1). (b) The HOMO picture of LaSi6 (C2v, 1A1). (c) The Mülliken spin density of YbSi6 (C2v, 2A1).

image file: c4ra03788j-f10.tif
Fig. 10 (a) The HOMO picture of LuSi6 (C5v, 1A′). (b) The SOMO picture of LuSi6 (C2v, 2A1). (c) The Mülliken spin density of LuSi6 (C2v, 2A1).

3.1. Geometric structure and stability

In this section, we discuss the structural details of a few low-lying isomers of MSi6q clusters, which were obtained by Saunders “Kick” (SK) global search technique combined with DFT geometric optimization (SK-DFT). We have obtained many low-lying isomers and determined the lowest-energy structures for RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) using the computation scheme SK-DFT described in Section 2 (computational methods). Here, in order to discuss the effects of doped impurity atom on pure silicon clusters, we carried out an additional SK-DFT calculation on pure Si6/7−,0 clusters. Many isomers are found and the top three isomers of Si6/7−,0 clusters are shown in Fig. 1. For Si6 cluster, the lowest-energy structure is isomer a, which can be viewed as a Si2-capping of a bent rhombus. The point group symmetry is C2v. The second lowest-energy isomer b is a regular octahedron with D4h symmetry, and it is only 0.19 eV above the ground state. Another structure d obtained for Si6 can be formed from capping of a trigonal bipyramid. It is less stable than isomer a by 0.85 eV. For neutral Si6 cluster, the lowest-energy isomer c is a relaxed octahedron. In fact, initial structures of isomers a and b are also considered in the present calculation. When six Si atoms were placed in the shape of isomer a or b, they move away from their original position during the geometry optimization process. The other two isomers, i.e., Si-capped triangular bipyramid-shaped isomer d and planar isomer e, are 0.62 and 1.45 eV higher than ground state structure c, respectively. The lowest-energy structures of Si6 and Si6 are in agreement with previously reported results by using different theoretical techniques.5,8,67 As for pure neutral and anionic Si7 clusters, three low-lying isomers (fh) of neutral structures are similar to the anionic structures, even the order of isomers is the same. The lowest-energy structure is a regular pentagonal bipyramid with D5h symmetry. Our calculated structure on neutral Si7 is in line with the early theoretical results.3,8,12 Isomers g and h are about 0.96, 0.97 eV (for neutral Si7) and 0.79, 0.87 eV (for anionic Si7), which is less stable compared to ground state structure f.

The ground-state structures and some low-lying metastable isomers of RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) species are shown in Fig. 1, in which the dark cyan spheres stand for silicon atoms and the black ones for RE atoms. Remarkably, our calculations show that all the low-lying isomers (top-eight) for MSi6q are three-dimensional (3D) structures. The spin multiplicity of ground state is kept constant for all other higher energy isomers (Table 1 and 2). It must be pointed out that the RE atom cannot be encapsulated inside the bare Si6q cluster and still remains on the surface of the lowest-energy MSi6q cluster using three different density functional methods. The most stable structure for RE-doped MSi6 (M = La, Ce, and Lu) is a 3D structure with the lanthanide metal atom sitting on top (or bottom) of the regular pentagonal bipyramid (see Fig. 1, isomer j). These most stable structures are formed by replacing a silicon atom on the top (bottom) of the inerratic pentagonal bipyramid shaped pure Si7 cluster. The other seven low-lying metastable 3D isomers are all much higher in energy than the perfect pentagonal bipyramid structure for all three clusters (0.29–1.44 eV for LaSi6, 0.34–1.35 eV for CeSi6 and 0.34–1.02 eV for LuSi6). The M–Si distance in the lowest-energy structures for RE-doped MSi6 (M = La, Ce, and Lu) has the following relationship: RLa–Si (2.96 Å) > RCe–Si (2.89 Å) > RLu–Si (2.82 Å), which is mainly due to the difference in atomic radius of lanthanide-metal RLa = 2.43 Å > RCe = 2.42 Å > RLu = 2.24 Å. It should be noted that all the ground state structures of MSi6 (M = La, Ce, and Lu) were retained after replacing a silicon atom on the top for the pure Si7 cluster, but electronic features had been significantly altered, which presented the reason for a far cry from photoelectron spectra of pure silicon clusters. On the other hand, results for the most stable structures of the RE-doped silicon clusters YbSi6 and MSi6 (M = La, Ce, Yb, and Lu) are also obtained. All the ground states of the neutral MSi6 (M = La, Ce, Yb, and Lu) and YbSi6 clusters also have a 3D structure, and no 2D isomers were obtained. The most stable structure for neutral MSi6 (M = La, Ce, Yb, and Lu) and YbSi6 is a 3D structure with the lanthanide metal atom sitting on side of the distorted pentagonal bipyramid (See Fig. 1, isomer i). For the neutral MSi6, the high symmetry 3D pentagonal bipyramid arrangement is not a stable structure, and is much higher in energy than the most stable distorted pentagonal bipyramid by 0.63, 0.62 0.90 and 0.51 eV for M = La, Ce, Yb, and Lu, respectively. As for YbSi6, the structure j is higher than i by 0.34 eV. According to the above descriptions of geometries, we find that all the lowest energy structures of neutral MSi6 (M = La, Ce, Yb, and Lu) and YbSi6 clusters were obtained by substituting one Si in the equatorial plane by an M atom from the ground state structure (Fig. 1, isomer f) of the Si7−,0 cluster; however, the anionic ground state of MSi6 (M = La, Ce, and Lu) clusters were obtained by substituting one Si by one M atom from the bottom (or top) of the pure Si7−,0 cluster.

It is interesting to know if the structure and stability of the clusters will change when an atom with a full 4f shell is doped into Si6 cluster compared with the case of by doping open 4f RE atoms. Here, we will compare the results of YbSi6 and LuSi6 with open 4f RE-doped MSi6 (M = Ce, Pr, Eu, Gd, and Ho) clusters.29,35 The point group symmetry of the lowest-energy structure of YbSi6 and LuSi6 is C2v symmetry. It should be mentioned that the early RE-doped MSi6 clusters have relatively low symmetry compared to the full 4f shell RE-doped clusters, e.g. CeSi6 (Cs) PrSi6 (Cs), GdSi6 (Cs), and HoSi6 (Cs). It seems that the more f-electrons, the higher the point group symmetry is. This may be attributed to Jahn–Teller distortion. For all the RE-doped MSi6 clusters, the additional RE atom always takes one of the equatorial positions of the pentagon, forming a distorted pentagonal bipyramid-shaped structure.

Table 1 gives various structural and energetic characteristics of both pure Si6/7q (q = 0, −1) and MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters. In order to analyze the stability of Re-doped clusters, we calculated the binding energies (BE) per atom and the doping energy (DE), which are defined by the following formulas: BE = [5E(Si) + E(Siq) + E(M) − E(MSi6q)]/7, DE = E(Si6q) + E(M) − E(M Si6q) (M = La, Ce, Yb and Lu; q = 0, −1), where E(Si), E(Siq), E(M), E(MSi6q) and E(Si6q) represent the total energies of the ground-state of Si, Siq, RE atom M, MSi6q and Si6q clusters, respectively. The BE per atom value of the lowest-energy MSi6q clusters is displayed in Table 1. The notable feature is that the BE per atom in the RE-doped MSi6q clusters is larger than that in the pure Si6/7q clusters, except for the case of YbSi6 and LuSi6 clusters. This suggests that the Si-M bond is stronger than the Si–Si bond, and the doped RE atom with open f shells can stabilize the pure pentagonal bipyramid structure. In fact, the high stability of these clusters has been observed in EuSi6/7/8.29 To further study the stabilities of MSi6q clusters, we will discuss the DE of the ground state structure. From Table 1, we can see that both neutral and anionic CeSi6 clusters have the highest DE of 16.73 and 16.93 eV among the clusters considered in the present study. Thus, CeSi60,− should be the most stable one among these clusters. The highest occupied molecular orbital-lowest unoccupied molecular orbital (HOMO–LUMO) energy gap is another useful quantity for examining the kinetic stability of clusters. A large energy gap corresponds to a high strength required to perturb the electronic structure. To calculate the HOMO–LUMO gap we consider HOMO and LUMO of both the spins (up and down). Comparing the difference between the energy gap of pure and doped clusters in Table 1 one can see that the Egap of the LuSi6 with 1.36 eV is the largest one among the doped clusters but still smaller than the pure Si6 and Si7 clusters. This suggests that LuSi6 is relatively more chemically stable than the neighboring clusters. Overall, after doping an RE atom into the silicon cluster, the chemical stability of RE-doped cluster is less reactive than pure cluster.

3.2. Comparison of the PES spectrum between experiment and theory

The well-resolved PES spectra for pure Si6/7 and doped MSi6 (M = Yb and Lu) in Fig. 2, 5 and 6 at 355 nm (3.496 eV), 266 nm (4.661 eV) and 213 nm (5.83 eV) photon energy are presented by the Castleman,4 Bowen46 and Nakajima47 groups, respectively, and they could serve as electronic fingerprints for the MSi6 clusters, and thus allow comparisons with theoretical calculations to verify the identified global minimum structures. The photo-electron spectra of present species are simulated by adding the occupied orbital energies relative to the HOMO (SOMO for open-shell systems) to the VDE and fitting them with a FWHM of 0.20 eV.53 The simulated photoelectron spectra for the two low-lying energy isomers of MSi6 (M = Yb and Lu) are presented in Fig. 5 and 6 for comparison. The low-lying energy isomers whose simulated spectra agree best with the experiment are assigned as the primary structure for most of the clusters. It must be pointed out that for MSi6 (M = La and Ce) clusters, to date no experimental PES has been reported; thus, to facilitate the future comparison with further experiments, we have drawn the simulated PES spectra of the four lowest-lying isomers within 0.40 eV at the PBEPBE/RE/SDD/Si/6–311+G(d) level in Fig. 3 and 4. Furthermore, experimentally measured ADE/VDE from the anion photoelectron spectroscopy are compared with the calculated values at the PBEPBE/RE/SDD/Si/6–311+G(d) level of theory in Table 3. The calculated first ADE/VDE energies of the host clusters Si6 and Si7 are also presented in Table 3 for comparison. The VDEs of each cluster anion corresponds to the first peak maximum of each spectrum in Fig. 2–6. As shown in Table 3, the ADE/VDE are 2.73/2.81, 2.66/2.83, 1.83/1.95, and 2.86/3.07 eV, respectively, for M = La, Ce, Yb and Lu, which are bigger than those of Si6/7 clusters (theoretical: 2.24/2.45 eV for Si6, 1.91/2.30 eV for Si7, experimental: the VDE is 2.39 eV for Si6, 1.85 ± 0.02/2.39 eV for Si7), except for the case of the YbSi6 cluster. The calculated ADE/VDE for pure Si6/7 agree with experimentally measured results.
3.2.1. Si6 and Si7. For pure Si6/7, the simulated PES spectra of the lowest-energy isomers a for Si6 and f for Si7 agree with the experimental spectra,4 which reproduce two prominent PES bands below 3.5 eV. The calculated first VDEs values are 2.45 and 2.30 eV, which are in agreement with the experimental values (2.39 ± 0.01 eV for Si6 and 2.39 ± 0.04 eV for Si7).4 The calculated ADE for pure Si7 is 1.91 eV, which is also in line with experimentally measured results (1.85 ± 0.02 eV).6 For Si6, the second lowest-energy isomer b is 0.19 eV higher in energy than a, and considering the inherent accuracy of the DFT, we could not exclude the probability of isomer b. However, the calculated first VDE (2.77 eV) is higher than the experimental value (2.39 ± 0.01 eV).4 According to the second criterion, we could exclude the possibility of the b isomer in the Si6 cluster beam. Thus, we can conclude that structures a and f should be the reasonable isomers for pure Si6/7. All other higher energy structures should have been negligible contribution to the PES spectrum. The very large X–A energy gap (1.10 eV for Si6, 1.50 eV for Si7) between the first and second VDEs indicates that the neutral Si6/7 clusters are closed-shell singlets with strongest stability.
3.2.2. LaSi6. The ground state of LaSi6 is closed shell (C5v, 1A′) with a valence configuration of (1a′)2(1a′′)2(2a′)2(3a′)2(2a′′)2(4a′)2(5a′)2(6a′)2(3a′′)2(4a′′)2(7a′)2(5a′′)2(8a′)2(9a′)2. The first VDE corresponds to electron detachment from the 9a′ HOMO, which is an antibonding σ orbital (Fig. 7(a)). The SOMO (14a) of the neutral cluster upon removing an electron from the anion HOMO is shown in Fig. 7(b) and its spin density (Fig. 7(c)) is mainly on the La atom (0.68). As shown in Fig. 3, the simulated PES patterns on the basis of the four isomers appear to be rather different, the one isomer j being a bit loose and consisting of four electronic states within 0–4.661 eV, while, the others isomers (k, l, and m) presented more congested situations and consisting of six or seven electronic states. VDE calculations for the four low-lying isomers of LaSi6 were performed using the PBEPBE/RE/SDD/Si/6–311+G(d) level of theory. The lowest-energy structure j gives the first ADE/VDE values (2.73/2.81 eV), which are bigger than the pure silicon clusters (2.24/2.45 eV for Si6, and 1.91/2.30 eV for Si7).
3.2.3. CeSi6. For CeSi6, to date, no experimental PES spectrum has been reported; thus, we only present our theoretical detachment spectrum from the lowest-energy and three others isomers of the CeSi6 cluster at the PBEPBE/RE/SDD/Si/6–311+G(d) level in Fig. 4. The simulated PES spectra and calculated ADE/VDE of the four low-lying isomers are presented in Fig. 4 and Table 3. The ground state of CeSi6 is open shell (2A′′) with a valence configuration of (1a′)2(1a′′)2(2a′)2(3a′)2(2a′′)2(4a′)2(5a′)2(3a′′)2(6a′)2(4a′′)2(7a′)2(5a′′)2(8a′)2(9a′)2(6a′′)1. The 6a′′ SOMO of CeSi6 (Fig. 8(a)) is primarily a nonbonding σ orbital. The Mülliken spin density (Fig. 8(d) and (e)) shows that the unpaired spin is primarily occupied on the Ce atom. After removal of an electron, the ground state structure of the neutral CeSi6 is slightly different from CeSi6, although it retains the Cs symmetry. The other low-lying isomers (k, l, and m) are found to be nearly degenerate in energy (ΔE = 0.34, 0.37 and 0.38 eV), and their simulated spectra are a lot different from the ground state structure. They reproduce much more prominent PES bands than the global minimum structure j. The first calculated VDE (2.83 eV), corresponding to the electron detachment from the 6a′′ SOMO is larger than three others isomers (2.18 eV for k, 2.19 eV for l and 2.08 eV for m, Table 3). Detachment from the fully occupied MOs below the 6a′′ SOMO can lead to both singlet (S) and triplet (T) final states. At the PBEPBE level, the calculated VDEs from the (9a′)2 (T, S), (8a′)2 (T, S), and (5a′′)2 (T, S) orbitals correspond to the intense and broad A band (ranging from 3.14 to 3.25 eV).
3.2.4. YbSi6. The C2v, (2A1) structure (Fig. 5 i) is clearly the ground state for YbSi6 with alternative structures being at least 0.34 eV higher in energy. The ground state of YbSi6 is open shell (2A1) with a valence configuration of (1a1)2(1b2)2(2a1)2(1b1)2(3a1)2(2b2)2(4a1)2(1a2)2(5a1)2(2b1)2(3b2)2(6a1)2(4b2)2(7a1)2(3b1)2(5b2)2(8a1)2(4b1)2(2a2)2(9a1)2(6b2)1. Experimental photoelectron spectra of YbSi6 generally differ from those of Ho-, Gd-, and Pr-doped LnSi6 members with their most characteristic distinction being filled comparably low EBE peaks.46 The X band in the experimental spectra is a low electron binding energy peak, which is broad and congested, suggesting that it may contain multiple detachment channels. Indeed, the calculated VDEs for the first seven detachment channels, from the SOMO 6b2 (1.95 eV), 9a1 (1.97 eV and 1.99 eV), 2a2 (2.02 eV and 2.04 eV) and the 4b1 (2.05 eV and 2.06 eV) orbital (T and S) are very close to each other, consistent with the broad X band. The 6b2 SOMO is mainly a nonbonding σ orbital, as shown in Fig. 9(a) and (c) for the spin density. On removal of an electron, the neutral YbSi6 retains the C2v structure with a shorter Yb–Si bond by about 0.10–0.13 Å. The calculated VDEs from the 8a1 orbital, 2.62 eV(S) and 2.64 eV(T), corresponding to the removal of the spin up and spin down electrons, are in good agreement with the observed VDE of the A band (2.60 ± 0.05 eV).46 The simulated PES spectrum of the second-lowest energy isomer j is different from the experimental features, suggesting structure j should not be a reasonable isomer. Furthermore, the theoretical first ADE/VDE (2.39/2.42 eV) values of the second-lowest energy isomer j agree very poorly with the experimental data (1.80 ± 0.05/1.98 eV).46 Overall, the simulated PES spectrum and calculated ADE/VDE from the global minimum of YbSi6 are in very good agreement with the experimental observations (Table 3 and Fig. 5), lending considerable credence to the global minimum identified.
3.2.5. LuSi6. The comparisons between experiment and theory for the ground state of LuSi6 are presented in Table 3 and Fig. 6. The ground state of LuSi6 (structure j) is closed shell (1A′) with a valence configuration of (1a′)2(1a′′)2(2a′)2(3a′)2(2a′′)2(4a′)2(5a′)2(6a′)2(3a′′)2(7a′)2(4a′′)2(8a′)2(5a′′)2(9a′)2. The first calculated VDE corresponds to electron detachment from the 9a′ HOMO, which is mainly a bonding σ orbital (Fig. 10(a)). The SOMO (7a1) of the neutral cluster on removing an electron from the anion HOMO is shown in Fig. 10(b), and its spin density (Fig. 10(c)) is seen to concentrate on the Lu atom. The simulated PES patterns on the basis of the two isomers (Fig. 6) appear to be rather different, i.e., the PES of the one isomer j is extremely loose and consists of only two bands within 0–4.661 eV, while, the other isomer i presented extremely congested situations and consisting of five bands. For the ground state structure of LuSi6, the calculated VDE of 3.07 eV is in good agreement with the experimental value of 3.20 eV (Table 3);47 however, the calculated ADE of 2.86 eV is larger than the experimental value of 2.10 eV. As seen from Table 3, this situation is ambiguous when considering the first ADE. We note that the second isomer i is 0.35 eV above the global minimum structure, but the computed ADE (2.00 eV) agrees very well with the experimental results (2.10 eV).47 Usually, it is rare to have such a high-energy isomer coexisting in the cluster beam experiments. Considering their very close ADE, it is possible that isomer i is a minor isomer. These results lead us to conclude that the global minimum structure j and minor isomer i might be coexisting in the LuSi6 cluster beam.

3.3. Electronic and magnetic properties

We performed natural population analysis (NPA) for the lowest-energy structures of MSi6q clusters, and the summarized natural charges and populations are given in Table 4. As seen from Table 4, charge always transfers from the RE atom to the Si atom for RE-doped MSi6q clusters. The transferred charges are from 0.47 e to 0.91 e for neutral MSi6 and by 0.10 e–0.33 e for anionic species, which indicates that the RE atom acts as an electron donor. Note that the Yb atom loses the largest charge (0.91 e) from its neighbouring Si atoms. The NPA of all the Si atoms bonding to RE atom is negative. The charge distribution is dependent on the symmetry of the cluster. The magnetic properties are one of the most interesting in cluster physics. The physicochemical property of size-selected cluster is quite different from that of single molecular or bulk material. Unfortunately, no Stern–Gerlach experiment has been carried out for present species up to now; however, the reliability of the present DFT (PBEPBE) was validated by performing calculations on the simulated PES spectra, the first ADE/VDE on pure Si6/7 and many other RE-doped MSi6 (M = Yb, Lu, Pr, Gd, and Ho) clusters. Magnetic moment (μB) of 4f, 5d, 6s, and 6p states for the RE atom, total magnetic moment (μB) of the RE atom, and total magnetic moment of the lowest-energy RE-doped MSi6q (M = La, Ce, Yb and Lu; q = 0, −1) clusters are presented in Table 5. As shown in Table 5, it can be seen that the total magnetic moment of the CeSi6 cluster has the largest value (2 μB) among the MSi6q species. It can also be seen that the total magnetic moments of the MSi6q clusters and the magnetic moments on the RE do not quench when the RE is encapsulated in a Si6 outer frame structure, which is consistent with the results of the EuSin.29 Interestingly, the overwhelming majority of total magnetic moments are contributed by RE atoms. These results seem to indicate that the majority of f electrons do not form chemical bonds between the RE atoms and their silicon cluster environment. To further understand the magnetic properties of MSi6q clusters, we performed a detailed analysis of the local magnetic moment of RE atom in MSi6q clusters. For free Yb and Lu atoms, the configurations of valence electrons are [Yb] 4f146s2 and [Lu] 4f145d16s2, which have closed 4f shells and primarily donate their 6s electrons to Si6/Si6, forming charge-transfer YbSi6/LuSi6 clusters with little distortions. However, for the open 4f shell of Ce dopant, the free Ce atom with the configuration of valence electrons 4f15d16s2, the 4f orbitals obtain electrons by 0.13 e for CeSi6 and the electron transfer is very small for CeSi6 from the Si6/Si6 clusters.

Certain trends of the magnetic moments of MSi6q in this work can be described as follows: (1) Table 5 shows that LaSi6, CeSi60,−, YbSi6 and LuSi6 are magnetic clusters, and the magnetic moments are 1, 2, 1, 1 and 1 μB, respectively. LaSi6, YbSi6 and LuSi6 species present a non-magnetic ground state. (2) For all the RE-doped MSi6q clusters, RE atoms carry most of the magnetic moments. The magnetism of the clusters considered in this study is mainly located on the RE atoms, and only very small magnetic moments are found on the parent clusters. This is quite consistent with the result of Eu-doped Si clusters.29 For CeSi6 clusters, Si6 contributes a small negative magnetic moment to the total moment, while for other RE-doped clusters, they carry a small positive magnetic moment. (3) The dopant RE atoms remain largely localized, and the atomic-like magnetism is maintained in the doped clusters. The pure Si6 cluster is usually non-magnetic, but when being doped with different RE atoms, it sometimes becomes a magnetic cluster.

4. Conclusions

In conclusion, we have investigated the structural, electronic, and magnetic properties of a series of RE-doped neutral and charged silicon clusters MSi6q (M = La, Ce, Yb and Lu; q = 0, −), using previous experimental anion photoelectron spectroscopy and density-functional calculations at the PBEPBE level. Stochastic search procedure by Saunders “kick” method combined with extensive DFT calculations are used for global minimum searches (SK-DFT). Optimized geometries of the ground-state and low-lying energy states of both the anion and neutral species are reported using SK-DFT and the results are compared with experimental photoelectron spectroscopy results from the literature. The global minimum search revealed that the MSi6 (M = La, Ce, and Lu) is a three-dimensional structure with the lanthanide-metal atom sitting on top (or bottom) of the regular pentagonal bipyramid, which is obtained by substituting one Si by M atom from the bottom (or top) of the pure Si7−,0 cluster. Instead, the neutral species and YbSi6 show the impurity as a four-coordinate atom in the equatorial plane of a pentagonal bipyramid, which is obtained by substituting one Si in the equatorial plane by an M atom from the ground state structure of the Si7 cluster. The CeSi6−,0 clusters have the largest binding energy and doping energy, indicating that the neutral and anionic CeSi6 clusters should be the most stable among MSi6q species. The study of magnetic properties indicates that the magnetic moment of the doped RE atoms remains largely localized and the atomic-like magnetism is maintained in the doped clusters. The pure Si6 cluster is usually non-magnetic, but when being doped with different RE atoms, it sometimes becomes a magnetic cluster. The current Si6 doping with different RE atoms suggest that a new class of pentagonal bipyramid clusters with varying magnetic properties may exist. The systematic methods, including structural characteristics and magnetic properties, presented in this study are useful for the analysis of the theoretical and experimental data and provide insight into more complicated cluster systems.

Acknowledgements

This work was supported by the Natural Science Foundation of Fujian Province of China (no. 2012J05005 and 2014J05006), the Programs for New Century Excellent Talents and Excellent Youth Talents in Fujian Province University (NCETFJ and EYTFJ, no. JA13009), and the Fundamental Research Funds for the Central Universities (JB-ZR1201).

Notes and references

  1. W. L. Brown, R. R. Freeman, K. Raghavachari and M. Schlueter, Science, 1987, 235, 860 CAS.
  2. J. T. Lyon, P. Gruene, A. Fielicke, G. Meijer, E. Janssens, P. Claes and P. Lievens, J. Am. Chem. Soc., 2009, 131, 1115 CrossRef CAS PubMed.
  3. M. Haertelt, J. T. Lyon, P. Claes, J. Haeck, P. Lievens and A. Fielicke, J. Chem. Phys., 2012, 136, 064301 CrossRef PubMed.
  4. S. J. Peppernick, K. D. D. Gunaratne, S. G. Sayres and A. W. Castleman Jr, J. Chem. Phys., 2010, 132, 044302 CrossRef PubMed.
  5. W. A. Tiznado, P. Fuentealba and J. V. Ortiz, J. Chem. Phys., 2005, 123, 144314 CrossRef PubMed.
  6. C. Xu, T. R. Taylor, G. R. Burton and D. M. Neumark, J. Chem. Phys., 1998, 108, 1395 CrossRef CAS PubMed.
  7. K.-H. Ho, A. A. Shvartsburg, B. Pan, Z. Y. Lu, C. Z. Wang, J. G. Wacker, J. L. Fye and M. F. Jarrold, Nature, 1998, 392, 582 CrossRef CAS PubMed.
  8. A. Tekin and B. Hartke, Phys. Chem. Chem. Phys., 2004, 6, 503 RSC.
  9. O. Lehtonen and D. Sundholm, Phys. Chem. Chem. Phys., 2006, 8, 4228 RSC.
  10. X. Chu, M. Yang and K. A. Jackson, J. Chem. Phys., 2011, 134, 234505 CrossRef PubMed.
  11. R. L. Zhou and B. C. Pan, J. Chem. Phys., 2008, 128, 234302 CrossRef CAS PubMed.
  12. C. Pouchan, D. Bégué and D. Y. Zhang, J. Chem. Phys., 2004, 121, 4628 CrossRef CAS PubMed.
  13. R. R. Hudgins, M. Imai, M. F. Jarrold and P. Dugourd, J. Chem. Phys., 1999, 111, 7865 CrossRef CAS PubMed.
  14. S. M. Beck, J. Chem. Phys., 1989, 90, 6306 CrossRef CAS PubMed.
  15. P. Claes, V. T. Ngan, M. Haertelt, J. T. Lyon, A. Fielicke, M. T. Nguyen, P. Lievens and E. Janssens, J. Chem. Phys., 2013, 138, 194301 CrossRef PubMed.
  16. J. T. Lau, K. Hirsch, Ph. Klar, A. Langenberg, F. Lofink, R. Richter, J. Rittmann, M. Vogel, V. Zamudio-Bayer, T. Möller and B. v. Issendorff, Phys. Rev. A, 2009, 79, 053201 CrossRef.
  17. A. Grubisic, H. Wang, Y. J. Ko and K. H. Bowen, J. Chem. Phys., 2008, 129, 054302 CrossRef PubMed.
  18. E. Janssens, P. Gruene, G. Meijer, L. Wöste, P. Lievens and A. Fielicke, Phys. Rev. Lett., 2007, 99, 063401 CrossRef.
  19. K. Koyasu, J. Atobe, M. Akutsu, M. Mitsui and A. Nakajima, J. Phys. Chem. A, 2007, 111, 42 CrossRef CAS PubMed.
  20. K. Koyasu, M. Akutsu, M. Mitsui and A. Nakajima, J. Am. Chem. Soc., 2005, 127, 4998 CrossRef CAS PubMed.
  21. M. Ohara, K. Koyasu, A. Nakajima and K. Kaya, Chem. Phys. Lett., 2003, 371, 490 CrossRef CAS.
  22. M. Ohara, K. Miyajima, A. Pramann, A. Nakajima and K. Kaya, J. Phys. Chem. A, 2002, 106, 3702 CrossRef CAS.
  23. H. Hiura, T. Miyazaki and T. Kanayama, Phys. Rev. Lett., 2001, 86, 1733 CrossRef CAS.
  24. N. M. Tam, T. B. Tai, V. T. Ngan and M. T. Nguyen, J. Phys. Chem. A, 2013, 117, 6867 CrossRef PubMed.
  25. S. Jaiswal, V. P. Babar and V. Kumar, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 88, 085412 CrossRef.
  26. L. Guo, X. Zheng, Z. Zeng and C. Zhang, Chem. Phys. Lett., 2012, 550, 134 CrossRef CAS PubMed.
  27. H. Cantera-López, L. C. Balbás and G. Borstel, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 075434 CrossRef.
  28. R. N. Zhao, J. G. Han, J. T. Bai, F. Y. Liu and L. S. Sheng, Chem. Phys., 2010, 372, 89 CrossRef CAS PubMed.
  29. G. F. Zhao, J. M. Sun, Y. Z. Gu and Y. X. Wang, J. Chem. Phys., 2009, 131, 114312 CrossRef PubMed.
  30. R. N. Zhao, Z. Y. Ren, P. Guo, J. T. Bai, C. H. Zhang and J. G. Han, J. Phys. Chem. A, 2006, 110, 4071 CrossRef CAS PubMed.
  31. P. Guo, Z. Y. Ren, A. P. Yang, J. G. Han, J. Bian and G. H. Wang, J. Phys. Chem. A, 2006, 110, 7453 CrossRef CAS PubMed.
  32. J. G. Han, Z. Y. Ren and B. Z. Lu, J. Phys. Chem. A, 2004, 108, 5100 CrossRef CAS.
  33. V. Kumar, A. K. Singh and Y. Kawazoe, Nano Lett., 2004, 4, 677 CrossRef CAS.
  34. J. Lu and S. Nagase, Phys. Rev. Lett., 2003, 90, 115506 CrossRef.
  35. H. F. Li, X. Y. Kuang and H. Q. Wang, Phys. Lett. A, 2011, 375, 2836 CrossRef CAS PubMed.
  36. R. N. Zhao, J. G. Han, J. T. Bai, F. Y. Liu and L. S. Sheng, Chem. Phys., 2010, 378, 82 CrossRef CAS PubMed.
  37. T. T. Cao, X. J. Feng, L. X. Zhao, X. Liang, Y. M. Lei and Y. H. Luo, Eur. Phys. J. D, 2008, 49, 343 CrossRef CAS.
  38. H. Bai, H. J. Zhai, S. D. Li and L. S. Wang, Phys. Chem. Chem. Phys., 2013, 15, 9646 RSC.
  39. W. L. Li, C. Romanescu, Z. A. Piazza and L. S. Wang, Phys. Chem. Chem. Phys., 2012, 14, 13663 RSC.
  40. T. R. Galeev, A. S. Ivanov, C. Romanescu, W. L. Li, K. V. Bozhenko, L. S. Wang and A. I. Boldyrev, Phys. Chem. Chem. Phys., 2011, 13, 8805 RSC.
  41. H. J. Zhai, W. J. Chen, X. Huang and L. S. Wang, RSC Adv., 2012, 2, 2707 RSC.
  42. H. T. Liu, Y. L. Wang, X. G. Xiong, P. D. Dau, Z. A. Piazza, D. L. Huang, C.-Q. Xu, J. Li and L. S. Wang, Chem. Sci., 2012, 3, 3286 RSC.
  43. R. Pal, L. M. Wang, W. Huang, L. S. Wang and X. C. Zeng, J. Am. Chem. Soc., 2009, 131, 3396 CrossRef CAS PubMed.
  44. H. J. Zhai, B. Wang, X. Huang and L. S. Wang, J. Phys. Chem. A, 2009, 113, 3866 CrossRef CAS PubMed.
  45. H. J. Zhai, B. B. Averkiev, D. Yu. Zubarev, L. S. Wang and A. I. Boldyrev, Angew. Chem., Int. Ed., 2007, 46, 4277 CrossRef CAS PubMed.
  46. A. Grubisic, Y. J. Ko, H. Wang and K. H. Bowen, J. Am. Chem. Soc., 2009, 131, 10783 CrossRef CAS PubMed.
  47. K. Koyasu, J. Atobe, S. Furuse and A. Nakajima, J. Chem. Phys., 2008, 129, 214301 CrossRef PubMed.
  48. M. Saunders, J. Comput. Chem., 2004, 25, 621 CrossRef CAS PubMed.
  49. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS.
  50. M. J. Frisch, et al., GAUSSIAN09, Revision C.01, Gaussian, Inc., Wallingford, CT, 2010 Search PubMed.
  51. H. F. Li and H. Q. Wang, Phys. Chem. Chem. Phys., 2014, 16, 244 RSC.
  52. H. Q. Wang, H. F. Li and X. Y. Kuang, Phys. Chem. Chem. Phys., 2012, 14, 5272 RSC.
  53. H. Q. Wang and H. F. Li, J. Chem. Phys., 2012, 137, 164304 CrossRef PubMed.
  54. H. Q. Wang and H. F. Li, Chem. Phys. Lett., 2012, 554, 231 CrossRef CAS PubMed.
  55. H. Q. Wang, H. F. Li, J. X. Wang and X. Y. Kuang, J. Mol. Model., 2012, 18, 2993 CrossRef CAS PubMed.
  56. H. Q. Wang and H. F. Li, Comput. Theor. Chem., 2013, 1006, 70 CrossRef CAS PubMed.
  57. H. Q. Wang, H. F. Li and L. X. Zheng, J. Magn. Magn. Mater., 2013, 344, 79 CrossRef PubMed.
  58. X. Cao and M. Dolg, J. Mol. Struct.: THEOCHEM, 2002, 581, 139 CrossRef CAS.
  59. M. Dolg, H. Stoll and H. Preuss, J. Chem. Phys., 1989, 90, 1730 CrossRef CAS PubMed.
  60. C. Paduani and P. Jena, J. Nano Res., 2012, 14, 1035 CrossRef.
  61. A. D. Becke, Phys. Rev. A, 1988, 38, 3098 CrossRef CAS.
  62. J. P. Perdew and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 45, 13244 CrossRef.
  63. A. D. Becke, J. Chem. Phys., 1993, 98, 5648 CrossRef CAS PubMed.
  64. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785 CrossRef CAS.
  65. B. Mielich, A. Savin, H. Stoll and H. Preuss, Chem. Phys. Lett., 1989, 157, 200 CrossRef.
  66. D. J. Tozer and N. C. Handy, J. Chem. Phys., 1998, 109, 10180 CrossRef CAS PubMed.
  67. S. F. Li and X. G. Gong, J. Chem. Phys., 2005, 122, 174311 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.