Dynamic motion of an Lu pair inside a C76(Td) cage

Juanyuan Haoa, Fengyu Lib, Hongjiang Lia, Xiaoyu Chena, Yuyan Zhanga, Zhongfang Chenb and Ce Hao*a
aState Key Laboratory of Fine Chemicals, Dalian University of Technology, Panjin, 124221, People's Republic of China. E-mail: haoce@dlut.edu.cn
bDepartment of Chemistry, Institute for Functional Nanomaterials, University of Puerto Rico, San Juan, PR 00931, USA

Received 12th December 2014 , Accepted 8th April 2015

First published on 8th April 2015


Abstract

Relativistic density functional theory (DFT) computations were performed to investigate the dynamic motion of an encapsulated Lu pair inside a C76(Td) cage. The results revealed that the lowest-energy configuration of Lu2@C76(Td) adopts C2 symmetry; four electrons are transferred to the outer carbon cage and the two encapsulated Lu atoms form a metal–metal single bond (with an electronic structure of Lu24+@C764−), and the good electron delocalization in the C764−(Td) cage partially contributes the thermodynamic preference of Lu2@C76(Td). The rather small barrier (3.2 kcal mol−1) for Lu2 atoms to hop from one stable site to another leads to flexible motion of the Lu pair inside the parent fullerene cage, and the Oh symmetrical motion trajectory of two Lu atoms is consistent with the STM image. The computed 13C NMR spectrum with this trajectory also agrees well with the experimental results.


Introduction

Endohedral metallofullerenes (EMFs) have rather unique properties unexpected in empty fullerenes.1–3 These properties lead to the wide applications of EMFs in various fields, such as materials science, energy, medicine, etc.4–11 Especially, the dynamic behavior of encapsulated atoms in the cage makes EMFs potential candidates for molecular and nano-devices.12 Not surprisingly, many experimental and theoretical studies have been carried out to investigate the dynamic motion of encapsulated atoms in EMFs.13 The scenario of La or Y atom moving inside C82 (with C2 symmetry) was firstly predicted by Andreoni and Curioni using ab initio molecular dynamics (MD) simulations.14 Nishibori et al. provided the first experimental evidence of the La atom movement inside C82 cage, their maximum entropy method (MEM)/Rietveld analysis using synchrotron powder diffraction data indicated that La atom has a giant bowl-shaped movement at room temperature.15 However, the DFT studies by Jin et al. revealed a conflicting picture,16 they suggested that La atom undergoes a boat-shaped motion with a small amplitude inside C82(C2v) fullerene cage, rather than the bowl or hemisphere model with a large amplitude obtained from MEM/Rietveld analysis.

M2@C80 (M = La, Ce) are the representatives of dimetallofullerenes. For La2@C80(Ih), two La atoms circulate three-dimensionally inside the C80(Ih) cage, which was supported by the two carbon signals in 13C NMR spectrum and one peak in 139La NMR spectrum in experiments.17 A deeper insight into the La pair motion was given by MEM/Rietveld analysis, which demonstrated that the two La atoms present a trajectory that connects the six-membered rings of C80(Ih), or a pentagonal-dodecahedron,18 as well as by DFT computations.19 For Ce2@C80(Ih), its 13C NMR spectrum indicated that the Ce atoms circulate randomly, as La atoms do in La2@C80(Ih).20 The other isomer of Ce2@C80, which has a D5h C80 outer cage, was synthesized by Yamada et al.,21 the 13C NMR analysis and DFT computations revealed that the Ce atoms circulate two-dimensionally along a band of 10 contiguous hexagons inside the C80(D5h) cage.

C76 fullerene has two structural isomers (D2 and Td) satisfying the isolated pentagon rule (IPR). The pristine C76(Td) isomer has not been synthesized yet due to its inherent open shell electronic structure. Excitingly, in 2010, Umemoto et al. synthesized and isolated Lu2@C76 for the first time.22 The UV-Vis-NIR, STM and 13C NMR chemical shift analysis disclosed that the entire Lu2@C76 molecule has Td symmetry, and the charge state of (Lu2)6+@C766−(Td) was suggested. They also speculated that the two encapsulated Lu atoms rapidly rotate along a rhombic dodecahedron inside the cage to maintain the high symmetry. However, in 2011, by means of systematic density functional theory (DFT) computations and statistical thermodynamic analysis, Yang et al.23 revealed that the inner two Lu atoms form a metal–metal single bond, and only transfer four electrons to the fullerene sphere, resulting in a formal valence state of (Lu2)4+@C764−(Td); they also suggested that the Lu2 dimer can hop rapidly between six equivalent configurations in the fullerene cage at room temperature, giving rise to a trajectory as a tetrahedron in C76(Td). Such a controversy caught our interest: what is the most stable configuration for Lu2@C76? What is its bonding nature? What is the energy barrier for the dynamic motion of Lu atoms? Is the motion scenario a rhombic dodecahedron or a tetrahedron?22

In this paper, by means of the relativistic DFT computations, we first investigated different isomers of Lu2@C76 and located the transition states for the Lu motion between the lowest-energy isomers. The bonding nature of the thermodynamically most favorable Lu2@C76(Td) isomer was analyzed by quantum theory of atoms in molecules (QTAIM) method. Our computations support Yang et al.'s recent electronic structure assignment (Lu24+@C764−). Then, by analyzing the partial trajectory of Lu atoms surrounding the molecular geometric center, we obtained the Oh symmetrical motion trajectory of two Lu atoms inside C76 cage. Furthermore, following the Lu atoms motion trajectory, we calculated the 13C NMR spectrum of Lu2@C76. Both the computed Lu motion trajectory and 13C NMR spectrum agree well with the experimental measurement.

Computational methods

ADF2008.01 program24–26 and DMol3 code27 were employed for the relativistic DFT calculations. The relativistic effects were taken into account by using the zero-order regular approximation (ZORA)28–30 basis sets in ADF2008.01, as well as all electron relativistic core treatment in DMol3.

By ADF2008.01 program, full geometry optimizations without symmetry constraints were carried out using the generalized gradient approximation (GGA)31 exchange–correlation functional BLYP32,33 with a triple-zeta polarized (TZP) basis set (referred to as GGA-BLYP/TZP). In addition, the frozen-core approximation up to the 1s orbital for C atoms and the 5p orbital for Lu atoms was used. By DMol3 code, geometry optimizations and transition state calculations were done at the GGA-BLYP/DNP level.

To disclose the bonding nature of Lu2@C76, we performed QTAIM analysis34,35 of the most stable Lu2@C76. The wavefunction was generated at BLYP/6-31G* ∼ SDD36 in Gaussian using the optimized geometries (BLYP/TZP in ADF). We also tested BLYP/6-31G* ∼ SARC37 and BP86/6-31G* ∼ SARC32,33,38 methods, and they gave very close results to that of BLYP/6-31G* ∼ SDD.

To better understand the inferred extraordinary electronic structure of Lu24+@C764−, we evaluated the electron delocalization (aromaticity) of the low-lying C764− and C766− isomers. Nucleus-independent chemical shifts (NICS, in ppm), a simple and efficient method to evaluate aromaticity,39,40 were computed at the cage centers of the optimized geometries of the empty cages with Gaussian 09 program.41

The 13C NMR spectrum was calculated using gauge-independent atomic orbital (GIAO) method42 at the GGA-BLYP/TZP theoretical level by using the ADF2008.01 program. The chemical shifts were first evaluated relative to C60, then were referenced to the carbon disulfide (CS2) (δ(C60) 143.15 ppm vs. CS2).43

Results and discussions

Searching for the lowest-energy isomer and its bonding nature

To clearly describe the symmetry of a C76(Td) cage, we take a triangle patch under a circum-spherical surface as a representative patch of the C76(Td) cage (Fig. 1). The area of this patch is equal to 1/24 of the total C76 surface. On this representative patch, 16 key points are located at five kinds of carbon atoms (points C1, C2, C3, C4, C5), at seven kinds of bond center (point b1, b2, b3, b4, b5, b6, b7), at one five-membered ring's center (point p), and at three six-membered rings' centers (point h1, h2, h3). Every key point has its own local symmetry. For example, point L is at the intersectional carbon atom of three six-membered rings with C3v local symmetry; point N is at the center of the six-membered ring with C3v local symmetry; point M is at the center of 6–6 bond with C2v local symmetry. We define point O as the geometric center of the carbon cage. Then the boundary of the patch can be characterized by a threefold axis OL, a two-fold axis OM and a three-fold axis ON. Note that all the independent symmetrical elements belonging to the Td point group can be found in the representative patch. The entire surface of the polyhedron can be encompassed by performing different elemental symmetry operations on the patch. Therefore, we can consider the patch LMN (shaded) as the smallest structural unit rather than the whole cage.
image file: c4ra16236f-f1.tif
Fig. 1 (a) The geometric structure of C76(Td) cage with the representative triangle LMN path; (b) all the key points on the patch: C1, b7 and h2 represent a three-fold axis, a two-fold axis, and a three-fold axis, respectively.

When the two Lu atoms are encapsulated in the cage, the system reduces from the original Td symmetry of C76 to a lower symmetry depending on the relative positions of the two Lu atoms inside the cage. Corresponding to the key points in the representative patch, full geometry optimizations were carried out for all 16 possible isomers, which resulted in five configurations, referred to as C2, C1, Cs, C3v and D2d according to the geometric symmetry. Table 1 summarizes the Lu–Lu distances (RLu–Lu), the shortest Lu–C distances (RLu–C), the relative energies and HOMO–LUMO gap energies for these five configurations of Lu2@C76. The results obtained by both ADF and DMol3 are presented.

Table 1 Key geometric parameters, relative energies and HOMO–LUMO gap energies for the five configurations of Lu2@C76. The results outside and inside the parentheses are from ADF and DMol3, respectively
Isomer RLu–Lu (Å) RLu–C (Å) Relative energy (kcal mol−1) Gap (eV)
C2 3.36 (3.38) 2.42 (2.40) 0.0 (0.0) 0.75 (0.80)
C1 3.40 (3.46) 2.48 (2.43) 1.8 (0.4) 0.74 (0.79)
Cs 3.45 (3.50) 2.49 (2.45) 5.4 (3.2) 0.70 (0.75)
C3v 3.47 (3.50) 2.46 (2.42) 8.1 (4.7) 0.63 (0.72)
D2d 3.36 (4.00) 2.44 (2.41) 25.5 (24.8) 0.49 (0.60)


ADF and DMol3 give the same order of relative energies and HOMO–LUMO gaps for the five isomers. The C2 configuration is of the lowest energy and largest HOMO–LUMO gap. The thermodynamic stability is followed by the C1 and Cs isomers. The more symmetrical isomers, C3v and D2d, have rather high relative energies and smaller HOMO–LUMO gaps.

Fig. 2 presents the top and side view of the C2 configuration, the lowest energy isomer of Lu2@C76. Though the RLu–Lu and RLu–C of the five isomers are pretty close (see Table 1), the C2 configuration has the smallest RLu–Lu and RLu–C (3.36 and 2.42 Å, respectively, by ADF; the corresponding values are 3.38 and 2.40 Å by DMol3).


image file: c4ra16236f-f2.tif
Fig. 2 Top (a) and side (b) views of the lowest-energy isomer (with C2 symmetry) of Lu2@C76. Lu atoms are green balls and C atoms are gray.

In order to confirm the reliability of the predicted relative stabilities, especially the frozen-core approximation for the basis sets, we carried out additional optimizations for the C2 and C3v isomers using three different function, namely GGA-BP86, PBE44 and PW91,45 together with the TZP all electron basis set using ADF package. The computational results summarized in Table 2 show that these function give rather similar optimized geometries and relative energies with those at GGA-BLYP/TZP. The RLu–Lu and RLu–C have almost no change, the biggest differences are within 0.03 Å. The relative energy difference between the two isomers is nearly the same, the C2 configuration is about 11 kcal mol−1 lower than that of the C3v configuration. Therefore, we can expect that the employed BLYP functional and the TZP basis set with frozen-core approximation can result in reasonable structures and energies of all the configurations.

Table 2 Computed key geometric parameters and relative energies of the C2 and C3v isomers with different function
Method Isomer RLu–Lu (Å) RLu–C (Å) Relative energy (kcal mol−1)
BP C2 3.37 2.40 0.0
C3v 3.49 2.44 11.5
PBE C2 3.38 2.40 0.0
C3v 3.50 2.44 10.6
PW91 C2 3.37 2.40 0.0
C3v 3.49 2.44 11.3


The key point of the controversy about the electronic structure assignment ((Lu2)6+@C766− by Umemoto et al. vs. (Lu2)4+@C764− by Yang et al.) is whether the two encapsulated Lu atoms form a single bond. To address this issue, we performed QTAIM analysis, which is a well-established method to analyze the topology of the electron density, and has been used for revealing and quantifying the bonding situation between the metal atoms and carbon cage in EMFs.46–53 The result of the above obtained lowest-energy isomer of Lu2@C76 was shown in Fig. 3. Herein we focus on the encapsulated Lu atom pair. Our computations showed that there is one Lu–Lu bond critical point (BCP). The small values of electron density ρbcp (0.058 a.u.) and Laplacian 2ρbcp (0.005 a.u.) suggest weak covalent interaction between the two Lu atoms. The covalent nature is further confirmed by the negative total energy density (−0.021 a.u.) and the small ratio of the kinetic energy density G to ρ (<1). The covalent bonding between the encapsulated Lu pairs leads to the formal valence state of Lu24+@C764−, instead of Lu26+@C766−.


image file: c4ra16236f-f3.tif
Fig. 3 Molecular graphs of the most stable Lu2@C76 (C and Lu atoms are black and white balls, respectively, the BCPs are the tiny spheres in red). Ring and cage CPs are omitted for clarity.

With the formal valence state of Lu24+@C764−, four electrons should be transferred from the encapsulated Lu pair to the outer C76 cage. What factor is stabilizing the C76(Td) tetranion? Noting that aromaticity is playing an important role to stabilize spherical clusters,54,55 we compared the relative energies and computed the NICS values at the structural centers of low-lying C764− and C766− isomers (Table 3). Our computed relative energies are in good agreement with those reported by Yang et al. Among the four structural isomers we considered, the hexanions of three isomers (with D2, C2v and Cs symmetry) have more negative NICS values, thus higher aromaticity, than the corresponding tetranions. The only exception is the Td isomer, its C764− has more negative NICS value than C766−, implying the stronger aromaticity of C764−(Td), which should be partially responsible for its higher stability (Table 3), and echoes gracefully the formal valence state of Lu24+@C764− instead of Lu26+@C766−.

Table 3 Computed relative energies (at B3LYP/6-31G*) and NICS values (at GIAO-B3LYP/6-31G*//B3LYP/6-31G*) of C764− and C766−isomers
C76 spiral ID PG PA C764− C766−
ΔE (kcal mol−1) NICS (ppm) ΔE (kcal mol−1) NICS (ppm)
19151 Td 0 0.00 −7.93 6.83 −5.79
19150 D2 0 37.87 −12.91 31.58 −18.21
19138 C2v 1 17.12 −4.96 4.55 −14.00
17490 Cs 2 20.74 −11.36 0.00 −20.10


The partial charges on the two Lu atoms obtained from our natural population analysis are both +1.5 |e|. The large charge transfer phenomena were also found in other EMFs.56,57 For example, Popov and Dunsch's comprehensive theoretical analyses found that the metal atom charges in M3N@C2n (M = Sc, Y, La) are in the range of +1.6 to +1.9 |e|, the electronic structure of (M3N)6+@C2n6− was assigned.57

The transition state and motion trajectory

In Umemoto et al.'s study,22 they speculated that Lu atoms in the C76 cage move rapidly in a rhombic dodecahedron trajectory to maintain the molecule's Td-symmetry. However, Yang et al.23 suggested that a tetrahedron trajectory. What is the true trajectory?

In order to compute the trajectory of Lu atoms, we located the molecular transition state between two optimized C2 configurations. The obtained transition state has a Cs symmetry, the two Lu atoms are separated by 3.75 Å, and the smallest distance between Lu and the carbon cage is 2.42 Å. Fig. 4a and b are the top and side views of the transition state structure, respectively. The nature of the transition state was characterized by the only imaginary frequency of 54.8i cm−1 in the DMol3 computations. The activation barrier is only 3.2 kcal mol−1, which indicates that the two Lu atoms can hop from one C2 point to another around the geometric center of carbon cage. On the basis of these data, we can get the partial trajectory that Lu atoms move inside the cage along the lowest energy path, as shown in Fig. 4c. We can further achieve the full motion trajectory by performing all the symmetry operations of Td group to the partial trajectory based on the representative patch of the C76(Td). The obtained trajectory is a rhombic dodecahedron with Oh symmetry (Fig. 4d). Note that eight apexes of the trajectory are small planes consisting of Lu atoms (pointing to h1 or C1 in the representative patch), and they look like truncated corners. The motion trajectory of two Lu atoms is consistent with the STM image in Umemoto et al.'s study.


image file: c4ra16236f-f4.tif
Fig. 4 Top (a) and side (b) views of the transition state structure; (c) the lowest energy path for the Lu atoms to hop between lowest-energy configurations, the blue balls represent the started C2 isomer, the green balls represent Cs isomer in transition state, the yellow balls represent the terminated C2 isomer, the arrow represent the motion direction; (d) the complete motion trajectory of Lu atoms in C76 carbon cage.

Why does the motion trajectory have Oh symmetry when it is operated in the Td group? There are two reasons:

(1). Not considering the outer C76(Td) carbon cage, the two Lu atoms are always moving around the center of symmetry i, which coincides with the geometric center O. By the symmetry operation:

S4i = C4, C3i = S6

All the symmetry operations of Td group associated with the center inversion will generate the Oh symmetry.

(2). When taking the carbon cage into consideration, the Lu2@C76 molecule will show Td symmetry due to the fast motion of Lu atoms inside the fullerene cage. As the Td group is a subgroup of Oh group, the motion trajectory of Lu atoms can be reduced from Oh to Td in the molecule. This trajectory is in good agreement with the experimental STM images.22

13C NMR spectrum

Utilizing the representative patch of the C76(Td), the 13C NMR spectra of C76(Td) and Lu2@C76 can be qualitatively discussed. There are five kinds of carbon atoms (points C1, C2, C3, C4, C5) in the patch, thus, 13C NMR spectrum should have five peaks. Points C3 and C4 should give full intensity signals because of their local C1 symmetry. In comparison, points C2 and C5 should present semi-intensity signals since they both have a mirror plane and local Cs symmetry, while with two mirror planes and a triple-axis, point C1 can only give 1/6 intensity signal due to its local C3v symmetry. Based on the above analysis, the 13C NMR spectra of C76(Td) and Lu2@C76 should have five peaks: two full intensity signals, two semi-intensity signals and one 1/6 intensity signal.

To simulate the dynamic 13C NMR spectrum of Lu2@C76 quantitatively, we computed the average chemical shifts for all of the carbon atoms and identified their status. According to the representative patch of the C76(Td), carbon atoms are classified and their arithmetic average chemical shifts are calculated. The computed 13C chemical shifts and intensities are 126.55 ppm (C1 intensity 1), 134.36 ppm (C3, intensity 6), 135.51 ppm (C2, intensity 3), 140.72 ppm (C5, intensity 3), and 141.01 ppm (C4, intensity 6), which agree very well with the experimental data (Table 4) and those predicted by Yang et al.23

Table 4 Comparison between the calculated and experimentally measured 13C NMR of Lu2@C76(Td) chemical shifts (δcal and δexp) and intensities (Ical and Iexp). The experimental results are from ref. 22
No. δcal δexp Ical Iexp C
1 126.55 129.61 1 1.00 C1
2 134.36 134.51 6 6.67 C3
3 135.51 137.85 3 3.18 C2
4 140.72 140.65 3 3.23 C5
5 141.01 142.39 6 6.55 C4


All the experimental data showed that Lu2@C76(Td) has Td symmetry,22 while our computations and those by Yang et al.23 showed that its lowest-energy isomer has C2 symmetry. How can we reconcile this discrepancy?

First we need to keep in mind that the experimentally measured UV-Vis-NIR, STM and 13C NMR chemical shifts present only the average result of the motion of Lu atoms in C76 cage. Actually, two stable sites for Lu atoms exist inside C76(Td) cage, which leads to the lowest-energy isomer with C2 symmetry. The small activation barrier (3.2 kcal mol−1) enables these two Lu atoms hop between the two energetically most preferred sites inside C76(Td). Thus, it is the dynamical motion behavior of Lu atoms along the rhombic dodecahedron trajectory (with Oh symmetry) that results in the overall Td symmetry of Lu2@C76.

The above conclusion that the dynamical motion of metal atoms inside the carbon cage can make the EMF molecule to maintain the symmetry of the parent fullerene cage is rather general. Similar situations were widespread in the other EMFs, and herein we give some examples. In M2@C80 (Ih, M = La, Ce) and Ce2@C80 (D5h), the motion of metal atoms makes the whole molecule to maintain Ih or D5h symmetry in the 13C NMR measurements.17,19–21,58,59 In M@C82 (C2v, M = La, Y, Sc, Gd), the motion of metal atoms makes the whole molecule to maintain C2v symmetry.13–16,60–62 In M@C74 (D3h, M = Ca, Y, La, Ba and Sr), the energy minimum has a C2v symmetry, while the 13C NMR showed D3h cage symmetry due to the motion of metal atoms.63–68 So far, all the experimental and theoretical results obey this rule. Finally, we propose that the overall symmetries of EMFs are determined by the symmetry of carbon cage, and the motion trajectory of metal atoms will act in concert with the cage to keep the symmetry.

Conclusions

By means of the relativistic DFT method and based on the representative patch of C76(Td), we obtained the lowest-energy isomer of Lu2@C76 (with a C2 symmetry). Our QTAIM analysis and the aromaticity evaluations showed that the formal electronic configuration of Lu24+@C764− rather than the traditional Lu26+@C766− can be inferred, which supports Yang et al.'s electronic structure assignment of Lu24+@C764−.23 The good electron delocalization in C764−(Td) partially contributes to the thermodynamic preference of Lu2@C76(Td). We further located the transition state structure for the Lu motion between the lowest-energy isomers, and obtained the motion trajectory of Lu atoms inside the C76 cage. The rather small transition barrier leads to the flexible motion of the encapsulated Lu atoms inside the C76 cage, and the computed Oh symmetrical motion trajectory of the two Lu atoms is consistent with the STM image. We also discussed the formation of this motion trajectory by group theory in detail. Additionally, the 13C NMR spectrum of Lu2@C76 quantitatively computed by this trajectory also agrees well with the experimental data. Our detailed analysis of the experimental and theoretical results indicate that that the overall symmetry of EMFs probably depends on the symmetry of carbon cage due to the dynamic motion of the encapsulated metal atoms.

Acknowledgements

This work has been supported in China by the National Natural Science Foundation of China (Grant nos 21036006, 21137001, and 21373042) and the State Key Laboratory of fine chemicals (Panjin) project (Grant no. JH2014009), and in USA by National Science Foundation (Grant EPS-1010094) and Department of Defense (Grant W911NF-12-1-0083). The computational resources were provided by Shanghai Supercomputer Center.

Notes and references

  1. T. Akasaka and S. Nagase, Endofullerenes: A New Family of Carbon Clusters[C], Springer, 2002 Search PubMed.
  2. H. Shinohara, Rep. Prog. Phys., 2000, 63, 843 CrossRef CAS.
  3. X. Lu, L. Echegoyen, A. Balch, S. Nagase and T. Akasaka, Endohedral Metallofullerenes, Fundamentals and Applications, CRC Press, 2014 Search PubMed.
  4. S. I. Kobayashi, S. Mori, S. Iida, H. Ando, T. Takenobu, Y. Taguchi, A. Fujiwara, A. Taninaka, H. Shinohara and Y. Iwasa, J. Am. Chem. Soc., 2003, 125, 8116 CrossRef CAS PubMed.
  5. T. Tsuchiya, R. Kumashiro, K. Tanigaki, Y. Matsunaga, M. O. Ishitsuka, T. Wakahara, Y. Maeda, Y. Takano, M. Aoyagi, T. Akasaka, M. T. H. Liu, T. Kato, K. Suenaga, J. S. Jeong, S. Iijima, F. Kimura, T. Kimura and S. Nagase, J. Am. Chem. Soc., 2007, 130, 450 CrossRef PubMed.
  6. R. B. Ross, C. M. Cardona, D. M. Guldi, S. G. Sankaranarayanan, M. O. Reese, N. Kopidakis, J. Peet, B. Walker, G. C. Bazan, E. V. Keuren, B. C. Holloway and M. Drees, Nat. Mater., 2009, 8, 208 CrossRef CAS PubMed.
  7. A. A. Popov, S. Yang and L. Dunsch, Chem. Rev., 2013, 113, 5989 CrossRef CAS PubMed.
  8. X. Lu, T. Akasaka and S. Nagase, Acc. Chem. Res., 2013, 46, 1627 CrossRef CAS PubMed.
  9. P. Jin, C. Tang and Z. Chen, Coord. Chem. Rev., 2014, 270–271, 89 CrossRef CAS PubMed.
  10. L. Dunsch and S. Yang, Small, 2007, 3, 1298 CrossRef CAS PubMed.
  11. L. Dunsch and S. Yang, Electrochem. Soc. Interface, 2006, 15, 34 CAS.
  12. J. Tang, G. Xing, Y. Zhao, L. Jing, H. Yuan, F. Zhao, X. Gao, H. Qian, R. Su, K. Ibrahim, W. Chu, L. Zhang and K. Tanigaki, J. Phys. Chem. B, 2007, 111, 11929 CrossRef CAS PubMed.
  13. M. S. Golden, T. Pichler and P. Rudolf, Struct. Bonding, 2004, 109, 201 CrossRef CAS.
  14. W. Andreoni and A. Curioni, Phys. Rev. Lett., 1996, 77, 834 CrossRef CAS.
  15. E. Nishibori, M. Takata, M. Sakata, H. Tanaka, M. Hasegawa and H. Shinohara, Chem. Phys. Lett., 2000, 330, 497 CrossRef CAS.
  16. P. Jin, C. Hao, S. Li, W. Mi, Z. Sun, J. Zhang and Q. Hou, J. Phys. Chem. A, 2007, 111, 167 CrossRef CAS PubMed.
  17. T. Akasaka, S. Nagase, K. W. M. Kobayashi, K. Yamamoto, H. Funasaka, M. Kako, T. Hoshino and T. Erata, Angew. Chem., Int. Ed. Engl., 1997, 36, 1643 CrossRef CAS PubMed.
  18. E. Nishibori, M. Takata, M. Sakata, A. Taninaka and H. Shinohara, Angew. Chem., Int. Ed., 2001, 40, 2998 CrossRef CAS.
  19. J. Zhang, C. Hao, S. Li, W. Mi and P. Jin, J. Phys. Chem. C, 2007, 111, 7862 CAS.
  20. M. Yamada, T. Nakahodo, T. Wakahara, T. Tsuchiya, Y. Maeda, T. Akasaka, M. Kako, K. Yoza, E. Horn, N. Mizorogi, K. Kobayashi and S. Nagase, J. Am. Chem. Soc., 2005, 127, 14570 CrossRef CAS PubMed.
  21. M. Yamada, N. Mizorogi, T. Tsuchiya, T. Akasaka and S. Nagase, Chem.–Eur. J., 2009, 15, 9486 CrossRef CAS PubMed.
  22. H. Umemoto, K. Ohashi, T. Inoue, N. Fukui, T. Sugai and H. Shinohara, Chem. Commun., 2010, 46, 5653 RSC.
  23. T. Yang, X. Zhao and E. Osawa, Chem.–Eur. J., 2011, 17, 10230 CrossRef CAS PubMed.
  24. E. J. Baerends, J. Autschbach, A. Bérces, C. B. P. M. Boerrigter, L. Cavallo, D. P. Chong, L. Deng, R. M. Dickson, D. E. Ellis, M. van Faassen, L. Fan, T. H. Fischer, C. F. Guerra, A. S. J. van Gisbergen, J. A. Groeneveld, O. V. Gritsenko, M. Grüning, F. E. Harris, P. van den Hoek, H. Jacobsen, L. Jensen, G. van Kessel, F. Kootstra, E. van Lenthe, D. A. McCormack, A. Michalak, V. P. Osinga, S. Patchkovskii, P. H. T. Philipsen, D. Post, C. C. Pye, W. Ravenek, P. Ros, P. R. T. Schipper, G. Schreckenbach, J. G. Snijders, M. Sola, M. Swart, D. Swerhone, G. te Velde, P. Vernooijs, L. Versluis, O. Visser, F. Wang, E. van Wezenbeek, G. Wiesenekker, S. K. Wolff, T. K. Woo, A. L. Yakovlev and T. Ziegler, ADF2008.01, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands, 2008 Search PubMed.
  25. G. te Velde, F. M. Bickelhaupt, E. J. Baerends, C. F. Guerra, S. J. A. V. Gisbergen, J. G. Snijders and T. Zifgler, J. Comput. Chem., 2001, 22, 931 CrossRef CAS PubMed.
  26. C. F. Guerra, J. G. Snijders, G. te Velde and E. J. Baerends, Theor. Chem. Acc., 1998, 99, 391 CAS.
  27. B. Delley, J. Chem. Phys., 1990, 92, 508 CrossRef CAS PubMed.
  28. E. V. Lenthe, R. V. Leeuwen, E. J. Baerends and J. G. Snijders, Int. J. Quantum Chem., 1996, 57, 281 CrossRef.
  29. E. V. Lenthe, J. G. Snijders and E. J. Baerends, J. Chem. Phys., 1996, 105, 6505 CrossRef PubMed.
  30. E. V. Lenthe, E. J. Baerends and J. G. Snijders, J. Chem. Phys., 1994, 101, 9783 CrossRef PubMed.
  31. S. H. Vosko, L. Wilk and M. Nusrir, Can. J. Phys., 1980, 58, 1200 CrossRef CAS PubMed.
  32. A. D. Becke, Phys. Rev. A: At., Mol., Opt. Phys., 1988, 38, 3098 CrossRef CAS.
  33. J. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 33, 8822 CrossRef.
  34. F. Biegler-König, AIM2000, version 1.0, University of Applied Sciences, Bielefeld, Germany, 2000 Search PubMed.
  35. R. F. W. Bader, Atoms in Molecules: A Quantum Theory, Oxford University Press, Oxford, U.K., 1990 Search PubMed.
  36. X. Cao and M. Dolg, J. Mol. Struct.: THEOCHEM, 2002, 581, 139 CrossRef CAS.
  37. D. A. Pantazis and F. Neese, J. Chem. Theory Comput., 2009, 5, 2229 CrossRef CAS.
  38. J. P. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 38, 7406 CrossRef.
  39. P. V. R. Schleyer, C. Maerker, A. Dransfeld, H. Jiao and N. J. R. V. E. Hommes, J. Am. Chem. Soc., 1996, 118, 6317 CrossRef CAS.
  40. Z. Chen, C. S. Wannere, C. Corminboeuf, R. Puchta and P. V. R. Schleyer, Chem. Rev., 2005, 105, 3842 CrossRef CAS PubMed.
  41. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, revision A. 01, Gaussian, Wallingford, CT, 2009 Search PubMed.
  42. K. Wolinski, J. F. Hinton and P. Pulay, J. Am. Chem. Soc., 1990, 112, 8251 CrossRef CAS.
  43. A. G. Avent, D. Dubois, A. Pénicaud and R. Taylor, J. Chem. Soc., Perkin Trans. 1, 1997, 2, 1907 RSC.
  44. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS.
  45. J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 46, 6671 CrossRef CAS.
  46. A. L. Svitova, K. B. Ghiassi, C. Schlesier, K. Junghans, Y. Zhang, M. M. Olmstead, A. L. Balch, L. Dunsch and A. A. Popov, Nat. Commun., 2014, 5, 3568 CAS.
  47. P. R. Varadwaj, I. Cukrowski and H. M. Marques, J. Phys. Chem. A, 2008, 112, 10657 CrossRef CAS PubMed.
  48. P. R. Varadwaj and H. M. Marques, Phys. Chem. Chem. Phys., 2010, 12, 2126 RSC.
  49. B. Tan, R. Peng, H. Li, B. Wang, B. Jin, S. Chu and X. Long, J. Mol. Model., 2011, 17, 275 CrossRef CAS PubMed.
  50. A. A. Popov and L. Dunsch, Chem.–Eur. J., 2009, 15, 9707 CrossRef CAS PubMed.
  51. P. R. Varadwaj and H. M. Marques, Theor. Chem. Acc., 2010, 127, 711 CrossRef CAS.
  52. P. R. Varadwaj, A. Varadwaj, G. H. Peslherbe and H. M. Marques, J. Phys. Chem. A, 2011, 115, 13180 CrossRef CAS PubMed.
  53. Z. Badri, C. Foroutan-Nejad and P. Rashidi-Ranjbar, Comput. Theor. Chem., 2013, 1009, 103 CrossRef CAS PubMed.
  54. M. Bühl and A. Hirsch, Chem. Rev., 2001, 101, 1153 CrossRef PubMed.
  55. Z. Chen and R. B. King, Chem. Rev., 2005, 105, 3613 CrossRef CAS PubMed.
  56. S. Osuna, M. Swart and M. Solà, Phys. Chem. Chem. Phys., 2011, 13, 3585 RSC.
  57. A. A. Popov and L. Dunsch, Chem.–Eur. J., 2009, 15, 9707 CrossRef CAS PubMed.
  58. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666 CrossRef CAS PubMed.
  59. Z. Huang, X. Liu, K. Li, D. Li, Y. Luo, H. Li, W. Song, L. Chen and Q. Meng, Electrochem. Commun., 2007, 9, 596 CrossRef CAS PubMed.
  60. E. Nishibori, S. Narioka, M. Takata, M. Sakata, T. Inoue and H. Shinohara, ChemPhysChem, 2006, 7, 345 CrossRef CAS PubMed.
  61. L. Liu, B. Gao, W. Chu, D. Chen, T. Hu, C. Wang, L. Dunsch, A. Marcelli, Y. Luo and Z. Wu, Chem. Commun., 2008, 474 RSC.
  62. S. Guha and K. Nakamoto, Coord. Chem. Rev., 2005, 249, 1111 CrossRef CAS PubMed.
  63. J. Xu, T. Tsuchiya, C. Hao, Z. Shi, T. Wakahara, W. Mi, Z. Gu and T. Akasaka, Chem. Phys. Lett., 2006, 419, 44 CrossRef CAS PubMed.
  64. H. Oliver, H. Martin, G. Arthur, M. Michael and D. P. J. Martin, Z. Anorg. Allg. Chem., 2005, 631, 126 CrossRef PubMed.
  65. S. Ren, D. Tian and C. Hao, J. Mol. Model., 2013, 19, 1591 CrossRef PubMed.
  66. T. Kodama, R. Fujii, Y. Miyake, S. Suzuki, H. Nishikawa, I. Ikemoto, K. Kikuchi and Y. Achiba, Chem. Phys. Lett., 2004, 399, 94 CrossRef CAS PubMed.
  67. W. Zheng, S. Ren, D. Tian and C. Hao, J. Mol. Model., 2013, 19, 4521 CrossRef CAS PubMed.
  68. S. Ren, D. Tian and C. Hao, Comput. Theor. Chem., 2013, 1020, 57 CrossRef PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.