A DFT based prediction of gold fullerene Au92Si12 with the aid of silicon

Seema Gautama, Neetu Goelb and Keya Dharamvir*a
aDepartment of Physics and Centre of Advanced Studies in Physics, Panjab University, Chandigarh-160014, India. E-mail: keya@pu.ac.in
bDepartment of Chemistry and Centre of Advanced Studies in Chemistry, Panjab University, Chandigarh-160014, India

Received 27th December 2013 , Accepted 28th January 2014

First published on 29th January 2014


Abstract

Structural evolution of the gold Bucky ball Au92Si12 has been studied via a systematic investigation of AunSi clusters within the framework of DFT with spin-polarized generalized gradient approximation (GGA). Beginning with low energy isomers of Aun (n = 1–16), we find that Si, being hypervalent, can attach onto the Au cluster in more than one valence configuration, leading to several possible geometrical arrangements of surrounding Au atoms. One such geometry has a Au–Si unit dangling over a quasi-planar arrangement of gold atoms. It is observed that a pentagonal unit of gold atoms attached with a Au–Si dangling unit can only exist in the presence of other neighbouring gold atoms to present a quasi-planar cluster. By repeating such quasi-planar clusters, a stable geometry of golden Bucky ball Au92Si12 with a binding energy per atom (Eb/n) of 3.814 eV and a HOMO–LUMO gap (Eg) of 0.6 eV has been optimized. The current work suggests the exciting possibility of a golden Bucky ball that has the same rotational symmetry as the C60 molecule with a much larger volume.


Introduction

At the nanoscale, gold clusters have attracted considerable attention owing to their unique electronic and optical properties, as well as for their potential medical applications. Biocompatible gold nanostructures have reportedly found possible applications in photothermal therapy, radiotherapy and imaging of cancerous tissue.1–11 Free Aun clusters exhibit a variety of fascinating geometrical structures, e.g. planar structures for n = 3 to 12, anionic cages for n = 16 to 18, bulk-fragment pyramidal structures for n = 19 to 23, possible tubular cages for n = 24 and 26 (ref. 11b) and a gold fullerene at n = 32. The tetrahedral isomers of Au16 and Au17 were the first to be experimentally12 confirmed as small gold cages and Au32 was first suggested to be a “24-carat golden fullerene.”13,14 The large empty space inside the cage clusters immediately suggested that they can be doped with a foreign atom to produce a new class of endohedral gold cages analogous to endohedral fullerenes, which are expected to have potential medical applications owing to their biocompatibility and large cage volume that may facilitate encapsulation of biomolecules and drugs.

It is well known from a large number of previous studies that impurity atoms strongly influence the geometric, electronic, and structural properties of doped gold clusters that are sensitive towards the nature of the dopant atom. Wang et al.15 studied a series of doped gold anion clusters, Au16M (M = Si, Ge, Sn), and found that the global minima were dominated by the strong M–Au interactions, reminiscent of the Au14M cluster. AunSi has a tetrahedral coordinated Si with an Au–Si dangling unit, resembling that of the larger Si doped gold cluster Au16Si, while AuxM (M = Ge, Sn) have quasi-planar structures based on square-pyramid motifs. With the exception of sulfur, p-electron (M = Al, Si, P) doped gold clusters yield nonplanar geometries for Au5M, while those doped with s-electrons (Na, Mg) yield planar geometries.16 Zhang et al.17 have systematically investigated the properties of doped clusters Au6M (Al, Si, P, S and Cl), with the exception of the argon atom, doping enhances the binding strength of these clusters in comparison to Au7. Wang et al.18 have systematically investigated the electronic and structural properties of the doped clusters Au16M (M = Ag, Zn, In) by photoelectron spectroscopy and theoretical calculations and they found that the dopant atom does not significantly perturb the electronic and atomic structure of Au16, but simply donates its valence electrons to the parent Au16 cage.

Doping of Aun clusters with Si is of particular interest and the Au–Si interface has been studied extensively owing to its importance in microelectronics. Sun et al.19 have examined the stability of Au16Si endohedral clusters and found that Si atoms prefer to bind to the exterior surface of the Au16 cage. Au16Si clusters may no longer be viewed as 20-electron closed-shell systems in the sense of the Jellium model and should rather be considered as 16-electron systems because four electrons are needed for the local Au–Si bonding. This is consistent with the fact that Group IV elements tends to make covalent bonds. Particularly for Si and Ge, Kiran et al.20 reported experimental and theoretical evidence that a series of Si–Au clusters [SiAun, n = 2–4] are structurally and electronically similar to SiHn. Walter et al.21 have performed DFT studies on the endohedral doping of the Au16 cage by Al or Si and observed that the O2 molecule adsorbed over the Au16Si cluster is activated to a superoxo-species.

Though several measurements and much theoretical research have examined the properties of AunSi clusters, there remain some open questions, such as how can silicon be incorporated into nano gold, what would be the effect of dopant Si atoms on the doped clusters with increase of cluster size, how the enhancement of doping concentration would influence the Au cage structure, and can we expect quasi-planar Au–Si clusters by increasing the doping concentration, which may lead to the design of new hybrid Au–Si nanostructures for potential applications in microelectronics, catalysis, and jewellery making.

With the objective to find answers to the above questions, in the current work we doped Aun gold clusters with silicon atoms starting from n = 1 to n = 16 and obtained a large number of geometric isomers. One geometrical arrangement that attracted our attention was a Au–Si dangling unit over a gold pentagon surrounded by neighbouring gold atoms, which had quasi-planar structure. These quasi-planar isomers are the genuine minima on the potential energy surface and eventually led us to a stable gold cage. The present study provides evidence of a stable Si doped gold nanocluster, Au92Si12, which may be designated as a gold fullerene with enhanced functionality and a large cage volume in comparison to C60.

Computational method

The geometry optimizations were carried out within the framework of DFT with spin-polarized generalized gradient approximation (GGA)22 implemented with Gaussian-03 code.23

The accuracy and reliability of the PW91 functional for small gold clusters has been verified in earlier studies.24 Previously reported25–27 global minima of Aun (n = 1–14) clusters, which are known to be planar geometries, were used for Si atom doping. The initial geometries were prepared by placing a Si atom as a cap on the possible sites of low energy 2D isomers of the Aun (n = 1–14) clusters. Beyond this, the initial geometries (for n = 15 and 16) were generated by extending the quasi-planar isomer of Au14Si by adding one and two gold atoms, respectively. We investigated a large number of isomers of AunSi clusters in the size range n = 1–13 and to the best of our knowledge the isomers reported here are the lowest energy ones. However, beyond n = 13, other low energy isomers in the size range studied here are reported in the literature. This work does not contradict them and focuses on a quasi-planar arrangement of Au atoms including either tetrahedral or square-planar local arrangements around the dopant Si atom. Although such structures are not the global minima, they are the genuine minima on the potential energy surface as confirmed by frequency analysis. The main objective of this study is to demonstrate the evolution of a golden Bucky ball of 92 Au atoms with the help of Si atoms via the formation of quasi-planar silicon doped gold clusters. The triplet-valence basis set 6-311++g(d) was used for the Si atoms, while a Stuttgart/Dresden SDD28 basis set was adopted for the Au atoms. Frequency analyses have also been performed at the same level of theory, PW91/SDD at default spin, to ensure that all the reported isomers are the genuine minima on the potential energy surface.

The geometry optimization of the predicted Au92Si12 golden cage was performed using DFT formalism as implemented in VASP.29 The total-energy calculation was carried out based on the plane-wave expansion method employing scalar relativistic ultra-soft pseudo potentials and the spin-polarized version of the generalized gradient approximation (GGA) for exchange and correlation. A simple cubic cell of 15 Å dimension with the γ point for the Brillouin-zone integration was considered for these calculations. The geometries are considered to be converged when the force on each ion becomes 0.01 eV Å−1 or less. The cut-off energy for the plane-wave basis set was 300 eV. The total-energy convergence was tested with respect to the plane-wave basis set size and simulation cell size, and the total energy was found to be accurate to within 1 meV. This cage molecule needs verification in future experiments as well as calculation at an enhanced level of accuracy to determine its useful properties.

Results and discussion

Optimized geometries of various isomers of AunSi

Based on the optimized lowest energy structures of pure gold clusters, we carried out an extensive lowest energy structure search on silicon doped medium sized gold clusters. The fully optimized geometries of the AunSi (n = 1–16) clusters, including some low-lying geometric isomers with higher energies, are depicted in Fig. 1.
image file: c3ra47999d-f1.tif
Fig. 1 The lowest energy structures of optimized isomers of AunSi clusters. The isomers are designated by na–nd, where n represents the number of Au atoms in the AunSi cluster and a–d rank the isomer in descending order of binding energy. ΔE represents the relative energy of an isomer with respect to the lowest energy isomer.

The excellent agreement of our calculated parameters for Au2 (bond length (BL): 2.55 Å and binding energy per atom (Eb/n): 1.09 eV per atom) as well as AuSi (bond length (BL): 2.28 Å and binding energy per atom (Eb/n): 1.64 eV per atom) with previously reported experimental and theoretical values30,31 established faith in the accuracy of the employed method and the reliability of the atomic pseudo-potentials used in this work.

Table 1 presents the binding energy per atom (Eb/n) and HOMO–LUMO gap (Eg) of pure gold (Aun) and silicon doped gold clusters AunSi. In all the optimized geometries of the AunSi clusters, silicon prefers to occupy the most highly coordinated position (see Fig. 1). This observation is concordant with the fact that Si is well known to exhibit hypervalency. Consequently, doping the Aun clusters with Si results in early 3D onset geometry relative to pure Aun clusters. It has been reported that small gold clusters doped with impurity elements possessing p-electrons adopt 3D structures due to sp3 hybridization,16 while doping with transition metals induces planar structures.32–35

Table 1 Binding energy per atom (Eb/n) and HOMO–LUMO gap (Eg) of pure gold clusters (Aun) and silicon doped gold clusters (AunSi) with PW91/SDD for Au and 6-311++G(d) for Si
Cluster Symmetry Energy (in a.u.) Eb/n (in eV) Eg (in eV)
Au2 C1 −271.68299 1.09 2.04
(Au2Si) 2a C2v −561.16878 1.94 1.81
2b Cs −561.13278 1.60 0.40
Au3 C2v −407.52958 1.13 2.10
(Au3Si) 3a Cs −697.06531 2.10 2.00
3b D3h −697.05971 2.06 1.25
3c C1 −697.03805 1.91 0.87
Au4 C2v −543.41941 1.45 0.96
(Au4Si) 4a Td −832.97586 2.27 2.02
4b C4v −832.96825 2.23 2.11
Au5 C2v −679.2997 1.59 0.87
(Au5Si) 5a Cs −968.8265 2.11 0.76
5b C4v −968.8057 2.02 0.54
Au6 D3h −815.2104 1.82 2.14
(Au6Si) 6a CS −1104.7222 2.18 1.63
6b C2v −1104.7192 2.16 1.41
Au7 Cs −951.0678 1.78 1.26
(Au7Si) 7a Cs −1240.5971 2.15 1.30
7b Cs −1240.58986 2.13 1.42
7c Cs −1240.58389 2.11 0.65
Au8 D4h −1086.96915 1.90 1.54
(Au8Si) 8a Cs −1376.48048 2.17 0.61
8b Cs −1376.46608 2.12 1.32
8c C4v −1376.45870 2.09 0.75
Au9 C2v −1222.8364 1.88 0.87
(Au9Si) 9a Cs −1512.36956 2.19 0.84
9b C1 −1512.3503 2.13 0.85
Au10 D2h −1358.73812 1.97 1.32
(Au10Si) 10a Cs −1648.26052 2.21 1.03
10b Cs −1648.25394 2.19 1.36
Au11 Cs −1494.6119 1.55 1.10
(Au11Si) 11a C1 −1784.1332 2.19 0.42
11b C1 −1784.1319 2.18 1.01
Au12 D3h −1630.51568 2.04 0.97
(Au12Si) 12a C1 −1920.02117 2.20 0.75
12b C1 −1920.01706 2.19 0.64
12c C2v −1919.98455(img freq. = 1) 2.12 0.47
Au13 C1 −1766.38596 2.02 0.30
(Au13Si) 13a Cs −2055.91299 2.22 0.92
13b C1 −2055.90439 2.20 0.63
Au14_a C2v −1902.29955 2.08 1.57
(Au14Si) 14a C1 −2191.81375 2.25 0.84
14b C1 −2191.80713 2.24 1.19
14c C1 −2191.79908 2.22 0.62
14d C1 −2191.79693 2.21 0.59
14e C1 −2191.75525(img freq. = 1) 2.14 0.92
Au15 C1 −2038.17735 2.09 0.77
(Au15Si) 15a C1 −2327.69106 2.23 1.13
15b C1 −2327.67568 2.21 0.72
Au16 Cs −2174.07615 2.13 1.25
(Au16Si) 16a Cs −2463.60072 2.28 1.06
16b Cs −2463.58702 2.25 0.79
16c C1 −2463.58280 2.25 0.87
16d C2 −2463.52569 2.16 0.84
Au16Si3 C2 −3042.51667 2.34 0.86


The hetro-diatomic Au–Si has a bond length of 2.27 Å, smaller than that of Au2 (2.56 Å) and in good agreement with the available experimental result (2.26 Å).36 The lowest energy isomer of the Au2Si cluster (2a) forms an isosceles triangle. The next highest-energy isomer, i.e. 2b, is a linear structure. The lowest energy structure of Au3Si, 3a, is a capped triangle. The two other planar isomers, 3b and 3c, are 0.15 eV and 0.74 eV higher in energy than the lowest energy isomer. The most stable geometry of Au4Si, 4a, has a tetrahedral configuration.37 The other isomer that has been optimized, 4b, has a square pyramidal structure and is 0.21 eV higher in energy than the tetrahedral 4a. The Au4Si cluster exhibits the highest binding energy among all the clusters. The lowest energy configuration of Au5Si favors a top capped square prism with Cs symmetry. The lowest energy isomers of Au6Si and Au7Si prefer to have a tetra-coordinated Si atom and penta-coordinated Si for the low lying geometric isomers with higher energy. In contrast to this, the lowest energy isomers of AunSi (n = 8–13) exhibit hexa-coordinated silicon and tetra-coordinated Si for the low lying geometric isomers. All the lowest energy isomers, 8a–13a, of AunSi (n = 8–13) have the dangling Au–Si unit over a curved sheet of gold pentagons and are regarded as quasi-planar isomers. This structural preference is owing to the hypervalent character of silicon atoms. It is noteworthy that the recurring unit, hexa-coordinated Si with a gold pentagon, is not stable by itself and requires gold atoms in the neighbourhood for its existence. This view point is supported by the Au6Si structures. In these clusters the gold pentagon with the Si–Au dangling unit could not be optimized as the lowest energy minima (the same is true for Au7Si cluster) due to the lack of the neighbouring Au atoms required for its stability. In the case of Au8Si, the structure 8a consisting of a gold pentagon with a Au–Si unit is the lowest energy structure in comparison to the other isomeric structures (8b and 8c, see Fig. 1). The preference for this geometrical pattern continues for all the larger sizes studied here. The higher stability of hypervalent isomers suggests that Au–Si interactions dominate the Au–Au interactions up to this size.

The caged structure of pure Aun clusters beyond the size n = 13 is well reported. The exo and endohedral doping of Si in the Au14 cage led to collapse of the gold cage with Si residing on the cluster surface (14a and 14b, see Fig. 1). Out of the five optimized geometries of Au14Si, 14a–14d are confirmed as stable isomers, while 14e is a metastable geometry with an imaginary frequency. Quasi-planarity is clearly perceived in the 14c isomer, which has a Au–Si dangling unit with a gold pentagon surrounded by the neighbouring gold atoms.

In the case of Au15Si, two optimized structures have been obtained. The input geometry for 15a is generated by the exohedral doping of Si in the Au15 cage. This input geometry was optimized to a distorted cage with Si residing over the surface of the cluster. For 15b, the input was generated by adding one gold atom to the bridging gold site of 14c and on optimization a stable structure was obtained (see Fig. 1). It also has a dangling Au–Si unit on the surface of a fourteen gold atom cluster.

Four optimized structures have been obtained for the Au16Si cluster. The structure 16c has a pentagonal unit of gold atoms with a Au–Si dangling unit surrounded by neighbouring gold atoms. Though this is one of the low energy isomers, the lowest energy structure has a distorted cage of Au16 with Si residing over the surface of the cage. It is pertinent to mention here that for AunSi clusters beyond size n = 13, caged structures are the lowest energy structures on the energy scale. Nonetheless, the quasi-planar Si doped AunSi (n = 1–16) structures are permissible with the Au–Si unit on the top of a pentagonal Au5 unit that needs to be supported by neighbouring gold atoms.

For n = 1–13, the lowest energy structures are quasi-planar isomers. This implies that Au–Si interactions are dominant over Au–Au interactions. However, beyond this size, i.e. for n = 14–16, it has been found that caged structures are the lowest energy structures in comparison to the quasi-planar isomers (low energy lying isomers). This implies that in large sized AunSi clusters (n = 14–16), Au–Au interactions dominate over the Au–Si interactions and the effect of the dopant atom decreases with increase in cluster size. In this context, it is interesting to see the effect of enhancing the number of silicon dopant atoms in the Aun clusters. For this purpose, two symmetrically bent tips of 16d were doped with two Si atoms and the input geometry was designated as Au16Si3_initial (see Fig. 2). The optimized geometry, i.e. Au16Si3_final, shows that the silicon atoms lift the two bent tips of 16d and instead of inducing a curve in the planar geometry of the pure gold cluster, the silicon atoms make the cluster more planar. It is encouraging to notice that with a larger number of silicon atoms, the binding energy per atom (Eb/n) increased (see Table 1). Since the amount of Si doped is crucial to the curvature induced in the gold clusters, we limit the dopant concentration up to one Si atom in 16 atoms of gold.


image file: c3ra47999d-f2.tif
Fig. 2 Initial and final geometry of the Au16Si3 cluster.

Following the trend of the lowest energy isomers of AunSi (n = 1–13), the low energy quasi planar isomers of AunSi (n = 14–16) also contain hexa-coordinated Si with a gold pentagon that is not stable by itself, as shown by stability analysis of the Au6Si structures. These quasi-planar isomers are one of the local minima.

It is thus established that the Au–Si dangling unit plays an important role in stabilising a quasi-planar sheet of gold atoms and the pentagonal unit of Au must have some neighbouring gold atoms. The geometry of this quasi-planar configuration is reminiscent of curvature introduced in a hexagonal graphene sheet by the presence of a pentagon. Moreover, when such pentagons are repeated regularly, the resultant structure is a closed cage, e.g., C60, C80, C120 etc. Encouraged by this observation we generated the coordinates of golden Bucky ball Au92Si12 (see Fig. 3) parallel to the structure of C60 (consistent with the above stability criteria) and on optimization we obtained a stable geometry with a binding energy per atom (Eb/n) of 3.814 eV and a HOMO–LUMO gap (Eg) of 0.6 eV. This gold Bucky ball has a beautiful symmetric structure. In exception to the fullerene, this ball has twelve dangling units, which further enhance the stability of the ball structure. Consequently we can say that our results establish that gold Bucky ball Au92Si12 is one of the low lying energy isomers, if not the global minimum. The size of the system is prohibitive in exploring other structural arrangements of Au92Si12 clusters that may have some gold packed atoms. However our stability analysis confirms that this Bucky ball Au92Si12 is thermodynamically stable and this ball obeys the same rotational symmetries as the C60 molecule (I5). Similar to C60, this molecule has no two pentagons sharing a bond. The diameter of the golden cage is 13.77 Å compared to 7.1 Å for C60. The Au atom on any dangling unit may provide a site for functionalization by a variety of organic or inorganic units.


image file: c3ra47999d-f3.tif
Fig. 3 Optimized geometry of the gold fullerene Au92Si12.

Energy and stability

Based on the lowest energy structures, the relative stabilities of AunSi and Aun (n = 1–16) are discussed in terms of binding energies per atom, which are expressed as:
image file: c3ra47999d-t1.tif

image file: c3ra47999d-t2.tif
where E(AunSi), E(Aun+1), E(Au) and E(Si) are the total energies of the ground state AunSi, Aun+1, Au and Si, respectively. The binding energies of the Aun+1 and AunSi clusters as a function of the total number of atoms present in the cluster are plotted in Fig. 4. This illustrates that as the cluster size grows, the contribution to Eb from Au–Si (Au–Au) interactions increases. The binding energy for AunSi is significantly higher than the binding energy of Aun clusters (see Table 1). The additional energy is due to stronger interactions between the Au and Si atoms. Although the Eb/n of the AunSi clusters are higher than those of Aun+1, the difference becomes smaller for larger sizes, indicating that the influence of impurity decreases as the number of gold atoms increases. It can be found that the Eb/n curves of the AunSi and Aun+1 clusters increase noticeably from n = 1 to n = 4, then the values show a smooth growing tendency in the cluster size range of n = 5–16.

image file: c3ra47999d-f4.tif
Fig. 4 The binding energies per atom (Eb/n) for Aun and AunSi clusters are plotted against cluster size.

The relative stability order in a series of clusters can be illustrated more emphatically through the second-order energy difference as a function of cluster size, as shown in Fig. 5. The second difference in energy for clusters may be calculated as,

Δ2E(n) = [E(n − 1) + E(n + 1) − 2E(n)]


image file: c3ra47999d-f5.tif
Fig. 5 Second-order difference of energies for AunSi and Aun clusters plotted against cluster size.

In cluster physics, the quantity Δ2E(n) is commonly known to represent the relative stability of a cluster of size n with respect to its neighbour. Also Δ2E(n) can be directly compared to the experimental relative abundance: the peaks in Δ2E(n) are known to coincide with the discontinuities in the mass spectra.31 In view of second-order difference of energies, positive values of Δ2E(n) mean that the clusters are particularly stable.

It is found that the stability patterns of Aun and AunSi clusters follow a trend of odd–even oscillations. Aun at even numbers and AunSi clusters at n = 4, 6, 10, and 14, correspond to a higher stability compared to the neighbour. Therefore we can conclude that even-numbered clusters are more stable than odd-numbered clusters. The particularly large positive values of Δ2E(n) for Au6 and Au4Si further establish their particular stability in the series of clusters.

The stabilities of these clusters were also analyzed by examining fragmentation energies. Here, we consider two fragmentation channels involving either a Au atom or a Si atom. The fragmentation energies for a neutral AunSi cluster are calculated as

Δ2E(AunSi) = E(AunSi) − [E(Aun−1Si) + E(Au)]

Fig. 6 shows the fragmentation energies for the AunSi clusters with respect to Au and Si atoms. It is found that Δ1 ≥ Δ2 for all n, thus implying that the interaction of Si atoms with the Aun clusters is energetically more favorable than Au atoms, which is in agreement with the lower bond length of Au–Si than Au–Au. It is noticed that the curves have obvious even–odd oscillations, as previously seen in the binding energy plot.


image file: c3ra47999d-f6.tif
Fig. 6 Size dependence of the fragmentation energy (eV) of AunSi (n = 1–16) clusters via loss of a Si atom and a Au atom.

The highest-occupied molecular orbital (HOMO) and the lowest-unoccupied molecular orbital (LUMO) gap is related to chemical reactivity, systems with larger HOMO–LUMO gaps being less reactive. The HOMO–LUMO gaps of the lowest energy structures of the AunSi and Aun clusters versus cluster size is plotted in Fig. 7. Both AunSi and Aun clusters follow the odd–even oscillatory trend. The particularly high HOMO–LUMO gaps for Au4Si and Au6 clusters establish the particular stability of these cluster sizes.


image file: c3ra47999d-f7.tif
Fig. 7 Size dependence of the HOMO–LUMO gap of AunSi (n = 1–16) clusters.

Conclusions

We have reported the structural evolution of gold Bucky ball Au92Si12 via formation of a quasi-planar arrangement of Aun−1 clusters with Au–Si dangling units. The structure, stability and effect of doping Si in gold Aun (n = 1–16) clusters is investigated by DFT calculations. Doping of Aun clusters by Si leads to an overall improvement in the binding of the Aun clusters, as the energies of Si doped gold (AunSi) clusters are lower than those of pure gold Aun clusters. In all the optimized geometries of AunSi (n > 7), a pentagon of gold atoms attached with a dangling Au–Si unit is present. Interestingly, this pentagonal unit with the Au–Si dangling unit cannot exist independently and requires the presence of neighbours to support quasi-planarity. We conclude that the Au–Si dangling unit on the planar sheet of Aun−1 atoms plays a major role in stabilizing the quasi-planar gold cluster, and this led to the idea of optimization of a closed Si doped nanocluster, Au92Si12, that may be designated as a gold fullerene. The theoretical prediction of a Au92Si12 Bucky ball made in the present study paves the way for researchers for its experimental realization and further theoretical investigation.

Acknowledgements

SG is thankful to the Council of Scientific and Industrial Research (CSIR, New Delhi) with award ref. no. 20-6/2009(i) EU-IV dated 30.12.2009 for the Senior research fellowship.

References

  1. J. Chen, D. Wang, J. Xi, L. Au, A. Siekkinen, A. Warsen, Z. Li, H. Zhang, Y. Xia and X. Li, Nano Lett., 2007, 7, 1318 CrossRef CAS PubMed.
  2. J. M. Robinson, T. Takizawa and D. D. Vandre, J. Microsc., 2000, 199, 163 CrossRef CAS.
  3. R. L. Whetten, J. T. Khoury, M. M. Alvarez, S. Murthy, I. Vezmar, Z. L. Wang, P. W. Stephens, C. L. Cleveland, W. D. Luedtke and U. Landman, Adv. Mater., 1996, 5, 428 CrossRef.
  4. R. P. Andres, T. Bein, M. Dorogi, S. Feng, J. I. Henderson, C. P. Kubiak, W. Mahoney, R. G. Osifchin and R. Reifenberger, Science, 1996, 272, 1323 CAS.
  5. C. A. Mirkin, R. L. Letsinger, R. C. Mucic and J. J. Storhoff, Nature, 1996, 382, 607 CrossRef CAS PubMed.
  6. A. P. Alivisatos, K. P. Johnsson, X. G. Peng, T. E. Wilson, C. J. Loweth, M. P. Bruchez and P. G. Schultz, Nature, 1996, 382, 607 CrossRef PubMed.
  7. R. P. Andres, J. D. Bielefeld, J. I. Henderson, D. B. Janes, V. R. Kolagunta, C. P. Kubiak, W. J. Mahoney and R. G. Osifchin, Science, 1996, 273, 1690 CrossRef CAS.
  8. I. L. Garzón, K. Michaelian, M. R. Beltrán, A. Posada-Amarillas, P. Ordejón, E. Artacho, D. Sánchez-Portal and J. M. Soler, Phys. Rev. Lett., 1998, 81, 1600 CrossRef.
  9. I. L. Garzón, C. Rovira, K. Michaelian, M. R. Beltrán, P. Ordejón, J. Junquera, D. Sánchez-Portal, E. Artacho and J. M. Soler, Phys. Rev. Lett., 2000, 85, 5250 CrossRef.
  10. K. Michaelian, N. Rendón and I. L. Garzón, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 60, 2000 CrossRef CAS.
  11. (a) T. G. Schaaff and R. L. Whetten, J. Phys. Chem. B, 2000, 104, 2630 CrossRef CAS; (b) Priyanka and K. Dharamvir, Phys. Chem. Chem. Phys., 2013, 15, 12340 RSC.
  12. (a) X. Xing, B. Yoon, U. Landman and J. H. Parks, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 165423 CrossRef; (b) L. Wang, S. Bulusu, H. Zhai, X. C. Zeng and L. S. Wang, Angew. Chem., Int. Ed., 2007, 46, 2915 CrossRef CAS PubMed.
  13. M. P. Johansson, D. Sundholm and J. Vaara, Angew. Chem., Int. Ed., 2004, 43, 2678 CrossRef CAS.
  14. X. Gu, M. Ji, S. H. Wei and X. G. Gong, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 70, 205401 CrossRef.
  15. R. Pal, L. M. Wang, W. Huang, L. S. Wang and X. C. Zeng, J. Am. Chem. Soc., 2009, 131, 3396 CrossRef CAS PubMed.
  16. C. Majumder, A. K. Kandalam and P. Jena, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 205437 CrossRef.
  17. M. Zhang, S. Chen, Q. M. Deng, L. M. He, L. N. Zhao and Y. H. Luo, Eur. Phys. J. D, 2010, 58, 117 CrossRef CAS.
  18. L.-M. Wang, R. Pal, W. Huang, X.-C. Zeng and L. Wang, J. Chem. Phys., 2009, 130, 051101 CrossRef PubMed.
  19. Q. Sun, Q. Wang, G. Chen and P. Jena, J. Chem. Phys., 2007, 127, 214706 CrossRef PubMed.
  20. B. Kiran, X. Li, H. Zhai, L. Cui and L. Wang, Angew. Chem., Int. Ed., 2004, 43, 2125 CrossRef CAS PubMed.
  21. M. Walter and H. Hakkinen, Phys. Chem. Chem. Phys., 2006, 8, 5407 RSC.
  22. J. P. Perdew, in Electronic Structure of Solids' 91, ed. P. Ziesche and H. Eschrig, Akademie Verlag, Berlin, 1991 Search PubMed.
  23. M. J. Frisch, G. W. Truks and H. B. Schlegelet, et al. Gaussian-03, Revision C.3, Gaussian, Inc., Pittsburgh, PA, 2003 Search PubMed.
  24. A. V. Walker, J. Chem. Phys., 2005, 122, 094310 CrossRef CAS PubMed.
  25. X.-B. Li, H.-Y. Wang, X.-D. Yang and Z.-H. Zhu, J. Chem. Phys., 2007, 126, 084505 CrossRef PubMed.
  26. B. Assadollahzadeh and P. Schwerdtfeger, J. Chem. Phys., 2009, 131, 064306 CrossRef PubMed.
  27. S. Bulusu and X. C. Zeng, J. Chem. Phys., 2006, 125, 154303 CrossRef PubMed.
  28. M. Dolg, U. Wedig, H. Stoll and H. Preuß, J. Chem. Phys., 1987, 86, 866 CrossRef CAS PubMed; P. Schwerdtfeger, M. Dolg, W. H. E. Schwarz, G. A. Bowmaker and P. D. W. Boyd, J. Chem. Phys., 1989, 91, 1762 CrossRef PubMed.
  29. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558 CrossRef CAS; G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 49, 14251 CrossRef; G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15 CrossRef; G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef.
  30. X. Li, B. Kiran and L. S. Wang, J. Phys. Chem. A, 2005, 109, 4366 CrossRef CAS PubMed.
  31. K. Clemenger, Phys. Rev. B: Condens. Matter Mater. Phys., 1985, 32, 1359 CrossRef CAS.
  32. H. Tanaka, S. Neukermans, E. Janssens, R. E. Silverans and P. Lievens, J. Am. Chem. Soc., 2003, 125, 2862 CrossRef CAS PubMed.
  33. L. Gagliardi, J. Am. Chem. Soc., 2003, 125, 7504 CrossRef CAS PubMed.
  34. M. Heinebrodt, N. Malinowski, F. Tast, W. Branz, I. M. L. Billas and T. P. Martin, J. Chem. Phys., 1999, 110, 9915 CrossRef CAS PubMed.
  35. D. W. Yuan, Y. Wang and Z. Zeng, J. Chem. Phys., 2005, 122, 114310 CrossRef CAS PubMed.
  36. J. J. Scherer, J. B. Paul, C. P. Collier and A. O'Keefe, J. Chem. Phys., 1995, 103, 9187 CrossRef CAS PubMed.
  37. C. Majumder, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 235409 CrossRef.

This journal is © The Royal Society of Chemistry 2014