Geometrical stabilities and electronic properties of Sin (n = 12–20) clusters with rare earth holmium impurity: a density functional investigation

Run-Ning Zhaoa and Ju-Guang Han*b
aInstitute of Applied Mathematics and Physics, Shanghai DianJi University, Shanghai 201306, People’s Republic of China
bNational Synchrotron Radiation Laboratory, University of Science and Technology of China, Hefei 230029, People’s Republic of China. E-mail: jghan@ustc.edu.cn; Fax: +86-551-65141078

Received 5th October 2014 , Accepted 3rd November 2014

First published on 3rd November 2014


Abstract

HoSin (n = 12–20) clusters with different spin states have been systematically investigated by using density functional theory with the generalized gradient approximation. The total energies, growth-pattern and equilibrium geometries as well as the APT charges of the HoSin (n = 12–20) clusters are calculated. The relative stabilities in terms of the calculated atomic averaged binding energies and fragmentation energies are discussed, revealing that the cake-like HoSin (n = 16, 18, 20) clusters have enhanced stabilities. Furthermore, the highest occupied molecular orbital – lowest unoccupied molecular orbital (HOMO–LUMO) gaps of the HoSin (n = 12–17) are above 1.55 eV while HoSi16 has the largest one (1.95 eV). Interestingly, the calculated dipole moments of the cake-like HoSin (n = 16, 18, 20) clusters are very small, corresponding to the global minima. According to the calculated APT charges of the Ho atom in the HoSin (n = 12–20) clusters, the contribution of charge-transfer to the stability of HoSin clusters is briefly discussed, manifesting that the charges in HoSin clusters transfer from the Si atoms to the Ho atom. Additionally, the optimized geometries show that the rare earth Ho atom is completely encapsulated into the centre of the Si frame at n = 15. This finding is in good agreement with the available experimental results.


1. Introduction

Silicon and germanium clusters have been studied extensively by using both theoretical and experimental techniques, as they may be employed not only as model systems for investigating localized effects in the condensed phase, but also as building blocks for developing new silicon based nanomaterials with tunable properties.1–18 However, pure silicon clusters are unsuitable as building blocks because they are chemically reactive due to the existence of dangling bonds. Transition metal atom-doped silicon clusters, on the other hand, may tend to form closed-shell electronic structures that are of higher stability than the pure silicon clusters and represent a new class of endohedral clusters encapsulating metal atoms3–18 and have important applications in the semiconductor industry. What’s more, the properties can be changed greatly with the aid of the addition of “impurity” atoms.4,12,15 In addition, the transition metal (TM) atoms can saturate the dangling bonds on the surface of the Si frame, which stabilizes the Si cages. Encouraged by these results, we believe that searching for suitable impurity atoms to obtain silicon-based nanomaterials with the desired properties is possible.

Rare earth lanthanide metal (Ln)-doped silicon clusters have attracted wide interest. This is because the Ln atoms could retain significant portions of their magnetic moments even when enclosed by a silicon cage due to their localized f-electrons. Experimentally, Ohara et al. firstly investigated the TbSin (n = 6–16) clusters by using photoelectron spectra (PES) and a chemical-probe method. The Tb-doped silicon clusters are relatively more stable with regard to photofragmentation than the bare silicon clusters of a similar size.2 The electronic properties of silicon clusters containing a transition or lanthanide metal atom, MSin (M = Sc, Ti, V, Y, Zr, Nb, Lu, Tb, Ho, Hf, and Ta), were investigated using anion photoelectron spectroscopy at 213 nm. In the case of the elements Sc, Y, Lu, Tb, and Ho, the threshold energy of electron detachment exhibits local maxima at n = 10 and 16. The electronic characteristics of MSin are closely related to their geometric and electronic structures.3 Grubisic et al. carried out the study of LnSin (n = 3–13, Ln = Ho, Gd, Pr, Sm, Eu, and Yb) clusters,5 and found that the dramatically increased adiabatic electron affinities of the LnSin clusters were attributed to their inherent electronic stabilization. Stimulated by these experimental observations, several computational investigations have been performed for Ln-doped silicon clusters including YbSin (n = 1–13),4,8,9 EuSin (n = 1–13), and LuSin (n = 1–12) clusters.6–10 The calculated results indicate that the Ln in the lowest-energy LnSin geometries occupied a gradual sinking site, and the site varies between the Ln surface-absorbed forms to the Ln-encapsulated forms with the size of the Sin atoms increasing. Moreover, the charge-transfer between the Ln atom and the Sin framework depends on different kinds of Ln and the cluster size. The stabilities for the specific-sized Ln-silicon clusters are enhanced after the Ln is doped into the Sin frame.

In a way, rare earth metals are special transition metals, possessing many of the properties of these elements (such as the optical and magnetic properties, and so on). Although there is some research on the properties of Ln-doped silicon clusters, only systematic theoretical investigations on the electronic properties and geometries of small-sized HoSin (n = 1–12, 16, 18)10 and HoSin (n = 1–12, 20) clusters have been reported so far; however, the calculated results can not provide the threshold size of the Ho atom being encapsulated into the silicon frame and do not allow exploration of the growth-patterns and the cake-like Ho-doped Si geometries.

In order to understand the properties of HoSin clusters and the critical size of the Ho encapsulated into the Si frame and to reveal the unusual size-dependent properties of the rare earth Ho-doped Sin clusters, we carry out a detailed study of HoSin (n = 12–20) clusters with relativistic effects being taken into account. The main objective of this research, therefore, is to provide a detailed investigation of the equilibrium geometries, charge-transfer properties, relative stabilities, fragmentation energies [D(n,n − 1)], atomic averaged binding energies [Eb(n)], and HOMO–LUMO gaps of HoSin (n = 14–20) clusters, which can provide instrumental help in the quest for this type of cluster-assembled material. However, it should be pointed out that the relativistic effect of the rare earth Ln is a challenging problem because of the complicated 4f electrons.

2. Computational details

The explicit treatment of all the electrons in a cluster (including the heavy metal Ho element) having a large number of atoms constitutes a demanding computational task. One of the best ways to surmount this difficulty is to make use of electron core potentials (ECP), also known as relativistic pseudopotentials,19 by means of which only the valence electrons are explicitly treated. ECP calculations can actually provide accurate results for both homo- and heteronuclear clusters and their various combinations as firmly proved by previous investigations on actinide elements and silicon/germanium clusters.20,21 Therefore, the combination of density functional theory (DFT) methods with ECP provides a feasible and accurate approach to the electronic structure study of the HoSin clusters as shown below.

Present calculations are done at the level of the DFT with the hybrid exchange and correlation (X3LYP) functional in combination with the 6-31G basis sets for Si atoms and the Large-core Stuttgart effective core potentials ECP56MHF ECP for describing the Ho atom,19 as implemented in the Gaussian 09 code.22 The various HoSin isomers, optimized geometries, and their harmonic vibrational frequencies are obtained. With the aim to examine all stable geometries as minima and to estimate zero-point energy corrections that are included into the calculated electronic energies, vibrational frequencies are calculated at the same level of theory.

In order to test the reliability of our calculations, the HoSi molecule is calculated at the X3LYP level with ECP56MHF ECP for the Ho atom and 6-31G for the silicon atom. The calculated bond length, frequency, and electronic state is 2.733 Å, 288.6 cm−1, and 4Σ, respectively; this is in good agreement with the previously calculated values of 2.838 and 218.8 cm−1 at the mPW3PBE level in combination with the 6-31G* basis sets for the Si atom and the Large-core Stuttgart effective core potentials (ECP28MWB) ECP for describing the Ho atom,16 as well as the values (2.773 Å, 267.3 cm−1, and 4Σ) calculated at the X3LYP level with ECP56MWB ECP for the Ho atom and the 6-31G basis sets for the silicon atom, as well as those values (2.766 Å, 265 cm−1) calculated at the QCISD level with the same basis sets. The Si–Si bond length, frequency, and the electronic state of Si2 at the X3LYP level with 6-31G basis sets is calculated to be 2.236 Å, 507.4 cm−1, and 3Σg, respectively, which is in good agreement with the experimental results (2.246 Å, 510.98 cm−1, and 3Σg)23 and the calculated values of Si2 of 2.352 Å, 445.7 cm−1, and 3Σg, respectively, at the B3LYP/LanL2DZ level.17,24,25 The calculated bond length, frequency, and electronic state for the Ho2 dimer at the X3LYP level with ECP56MHF ECP for the Ho atoms is 3.239 Å, 110.7 cm−1, and 5Σg, respectively, which matches the calculated bond length of the Ho2 dimer (3.875 Å),26 however, it is not in agreement with a previous result of 4.727 Å;27 however, the values are in good agreement with the values (3.048 Å, 127.4 cm−1, and 5Σg) calculated at the X3LYP level with ECP56MWB ECP for the Ho atoms as well as the values (3.078 Å, 131.7 cm−1) at the QCISD level with ECP56MWB ECP basis sets. Consequently, the X3LYP method with ECP56MHF for the Ho atom and 6-31G basis sets for the Si atom is reliable enough to be applied to describe the properties of the HoSin clusters.

As far as the Ho-doped silicon clusters are concerned,5,10 the previously calculated results, without considering relativistic effects, can not match available experimental results and find the critical size of Ho encapsulated within silicon clusters. In order to explore the accurate and reliable critical size of Ho completely encapsulated into silicon clusters as well as the physical and chemical properties of the HoSin (n = 12–20) clusters, the systematic investigations on the different sized Ho-doped Sin clusters are carried out with relativistic effects being taken into account. Equilibrium geometries of the HoSin (n = 12–20) clusters and their relative stabilities are investigated, and their frequencies are examined. If the imaginary frequencies are found, a relaxation along the coordinates of the found imaginary vibrational mode will be carried out until the local minimum is actually obtained. In addition, the spin electronic state for each initial geometry is considered at least from spin S = 1/2, 3/2, 5/2, to 7/2, and to 9/2.

3. Results and discussion

3.1. Geometries and stabilities

HoSi12. Different from the transition metal (TM)-doped Si12 (TM = Re, W, Pd, and Zr) clusters,12,15,24,25 the deformed hexagonal prism, D6h 12a and 12b isomers, with Ho being capped onto the Si12 structure, is generated and its total energy is much lower than that of the other Ho-doped Si12 structures (Fig. 1). As seen from the examined equilibrium geometries of the HoSi12 clusters, the identified structures for the most stable HoSi12 are usually different from the ReSi12, ZrSi12, and WSi12 clusters. The two cluster series show different growth patterns in that the critical sizes for the TMSin (TM = Re, Zr, and W) endohedral geometries are specified to be n = 10 or 12 while the critical size for the Ho atom in the HoSin clusters being completely encapsulated into the caged silicon frame turns out to not be n = 12. The 12c and 12d isomers deviate to the tetragonal and pentagonal prisms with the Ho atom being localized at the surface site of the silicon frameworks; their total energies are higher than the most stable 12a isomer (Table 1). The 12e isomer with the Ho atom acting as one of the silicon atoms is the highest one in terms of total energy. One concludes that the geometries at the size of n = 12 are dominated by the Ho surface-capped structures.
image file: c4ra11828f-f1.tif
Fig. 1 The stable HoSin (n = 12–20) geometries.
Table 1 Symmetries (sym), APT, dipole moment (Dip), the HOMO and LUMO gaps (Egap), the lowest frequencies (Freq), spin (S), total energy (ET), relative energy (ΔE), and S2 operator of HoSin (n = 12–20) clusters
System Isomer APT S S2 Dip (Debye) Freq (cm−1) Egap (eV) ET (Hartree) ΔE (eV)
HoSi12 12a −0.16 1/2 0.79 0.61 35.4 1.782 −3509.429974 0.00
12b   1/2 0.77   73.1   −3509.425637 0.12
12c   1/2 0.767   56.3   −3509.424052 0.16
12d   1/2 0.77   61.9   −3509.416065 0.38
12e   1/2 0.79   30.2   −3509.415085 0.41
HoSi13 13a 0.01 1/2 0.78 1.45 73.8 1.866 −3798.873102 0.00
13b   1/2 0.78   55.6   −3798.854091 0.52
13c   1/2 0.77   41.5   −3798.850846 0.61
13d   1/2 0.77   54.1   −3798.850604 0.612
HoSi14 14a −0.77 1/2 0.79 1.08 50.9 1.81 −4088.305069 0.00
14b   1/2 0.79   6.3   −4088.301805 0.09
14c   1/2 0.78   45.8   −4088.295465 0.26
14d   1/2 0.78   52.5   −4088.288696 0.45
14e   1/2 0.76   64.2   −4088.301875 0.086
HoSi15 15a −1.12 1/2 0.78 1.62 39.2 1.67 −4377.763301 0.00
15b   1/2 0.77   44.2 1.58 −4377.736937 0.72
15c   1/2 0.78   35.7   −4377.732548 0.84
15d   1/2 0.76   35.6   −4377.730276 0.90
HoSi16 16a −1.78 1/2 0.79 0.03 36.3 1.95 −4667.229364 0.00
16b   1/2 0.77   37.4   −4667.197371 0.87
16c   1/2 0.78   56.5   −4667.193831 0.97
16d   1/2 0.77   62.0   −4667.190793 1.05
HoSi17 17a −1.20 1/2 0.79 0.71 62.0 1.55 −4956.651556 0.00
17b   1/2 0.78   55.4   −4956.648025 0.096
17c   1/2 0.77   27.4   −4956.606087 1.237
17d   1/2 0.76   41.7   −4956.597174 1.48
HoSi18 18a −1.09 1/2 0.77 0.31 35.3 1.30 −5246.133360 0.00
18b   1/2 0.77   71.8   −5246.113030 0.553
18c   1/2 0.78   42.9   −5246.085834 1.293
18d   1/2 0.76   50.1   −5246.070841 1.701
18e   1/2 0.77   56.8   −5246.116888 0.45
HoSi19 19a −1.42 1/2 0.78 1.34 43.3 1.12 −5535.567366 0.00
19b   1/2 0.78   46.5   −5535.473914 2.54
19c   1/2 0.76   43.7   −5535.469922 2.65
19d   1/2 0.77   36.6   −5535.463601 2.82
HoSi20 20a −0.94 1/2 0.77 0.00 41.6 1.45 −5825.056184 0.00
20b   1/2 0.78   32.3   −5824.933468 3.34
20c   1/2 0.77   29.9   −5824.911108 3.95


HoSi13. The 13a geometry is generated after one silicon atom is added onto the 12c isomer. Its total energy is the lowest one. As seen from Fig. 1, it is obvious that the Ho atom still localizes at the surface site of the silicon frame and interacts with the silicon atoms. Furthermore, 13a forms a cage-like geometry, and the Ho atom in 13a does not tend to be moved into the centre of the silicon frame. On the contrary, the 13b geometry with the Ho atom being localized at the centre of the deformed hexagonal prism is higher in energy than 13a. It is generated from 12d. As for the 13c and 13d isomers, the Ho atom acts as a surface silicon atom and the low-lying geometries are formed.
HoSi14. n = 14 is the critical size for Zr atom-doped silicon clusters. The Zr atom is completely encapsulated into the centre of the silicon frame. As can be seen from the findings related to the geometries of the examined systems of HoSi14, the 14a and 14b isomers are seen with the Ho atom being surface-capped on the irregular and seriously distorted hexagonal prism Si12 frame; these geometries are yielded after a silicon atom is added to the surface of 13b. As for the 14c isomer, it can be described as the Ho atom being encapsulated into the centre site of the distorted heptagonal prism D7h Si14 frame with the Ho atom interacting with all silicon atoms; its total energy is slightly higher than that of the 14a and 14b isomers. The 14d geometry with a lower stability and higher energy deviates to the 14a and 14b geometries. The Ho atom in the 14d geometry acts as a surface silicon atom and contributes to form a pentagonal prism. Generally, the Ho atom in the 14a geometry shows a tendency to be moved into the silicon frame. On the basis of this finding, one can predict that the Ho atom is squeezed out of the silicon cage when it is small because Ho needs a larger cage in order to be situated within it.
HoSi15. The threshold size of the endohedral HoSin clusters turns out to be n = 15. The observed 15a structure, which can be seen as evidence of the Ho-encapsulated silicon framework, is the most stable structure; in comparison with the 14a geometry, the Ho in the 15a isomer, which can be described as the Ho atom interacting with silicon atoms directly, moves gradually into the centre site of two hexagons of the silicon framework; this geometry being generated after a silicon atom is added to the surface in the 14a frame is different from the most stable structures of the TM-doped Sin (TM = Mo and Zr) clusters.16,24,28 However, the other caged geometries 15b, 15c, and 15d, with the Ho atom being completely enclosed by silicon frames, have some dangling bonds being terminated by the enclosed Ho atom. The 15b and 15b geometries are obtained from the optimized C5v geometry and 15d is yielded after a silicon atom is surface-capped onto the 14c isomer. However, the stability of the irregular basket-like 15b and 15c structures along with the ring-like 15d structure is quite weakened as compared to that of 15a. Additionally, the Ho encapsulated within the heptagonal prism is an important block which is appropriate for the building block of quasi-one-dimensional Ho-doped silicon semiconductive nanomaterial. Furthermore, the critical size of the Ho atom-doped silicon clusters is larger than that of the TM atom-doped silicon clusters.29,30

Nakajima et al. investigated the adsorption reactivity of anionic HoSin and the experimental results suggest that the Ho is still incompletely encapsulated into the Si frame when the size of the Si cage swells to 16 atoms.3 One year later, photoelectron spectroscopy was utilized to study a variety of LnSin cluster anions (Ln = Yb, Eu, Sm, Gd, Ho, Pr; 3 ≤ n ≤ 13); for larger clusters resembling the Ho category it is predicted that the critical size of silicon clusters encapsulating Ho is localized at n > 11, approximately at about n = 11–15;5 interestingly, this suggestion is in good agreement with our calculated results of critical size n = 15. Moreover, the XLYP method in combination with effective ECP can give reasonable results of bond length and frequency. Consequently, our calculated results provide support for our selected methods.

HoSi16. For the HoSi16 clusters, the most perfect fullerene-like stable 16a isomer and the amorphous low-lying 16b and 16c isomers are optimized. The fullerene-like HoSi16 16a geometry with unbalanced Si bonds is different from that of the TM@Ge16 isomer17 and the bimetal Pd2-doped silicon frame.25 Additionally, another 16b isomer, which is obtained after the Si atom is capped onto the pentagonal prism 15a isomer, is higher in total energy than the fullerene-like 16a structure with a calculated frequency of 36.3 cm−1. The 16b isomer can be described as the Ho atom being encapsulated by two connected Si10 units. Furthermore, the 16b isomer is a new building block for forming new silicon clusters encapsulating Ho. After a Si atom is added to the surface of the 15d isomer, the sandwich-like octagonal prism 16c with Ho being localized within the silicon frame is obtained. Different from the 15d isomer in which Ho is encapsulated, the Ho atom in 16c isomer is encapsulated within the centre of an octagonal prism and interacts with 16 silicon atoms with inequivalent bond lengths, and tries to saturate the dangling bonds of the silicon atoms. As far as the 16d isomer is concerned, its geometry can be described as two silicon atoms being separately surface-doped on the deformed heptagonal prism HoSi14 frame; the 16d isomer is higher in total energy than the 16a isomer by 1.05 eV, reflecting the fact that the deformed octagonal prism 16c is more stable than the 16d isomer. As a consequence, the layered sandwich-like 16c and 16d isomers are not the dominant geometries because they are less stable than the cage-like 16a geometry.
HoSi17. With respect to the HoSi17 geometries, all optimized HoSi17 structures are shown as the sealed Si17 cage encapsulating Ho. The most stable fullerene-like HoSi17 17a isomer is generated, which can be described as one silicon atom being capped onto the 16a isomer. A low-lying 17b isomer can be found as a stable structure and its total energy is higher than that of the 17a isomer by 0.096 eV. As shown in Fig. 1, when one silicon atom is capped between the 1st and 7th silicon atoms of the HoSi16 16a isomer, a sealed caged 17b HoSi17 cluster is obtained. Another stable 17c isomer with the total energy being larger than the 17b isomer can be described as the Si-capped tetragonal prism and it is different from the 17b isomer in that the Si–Si dimer is symmetrically distributed at each side of the tetragonal prism. Additionally, one silicon capped octagonal prism 17d structure is found as the stable structure; however, its total energy is higher than that of the other isomers. This geometry is yielded from 16c. The 17d isomer is less stable than the 17a isomer because the Ho-coordinated silicon atoms in the former are bigger than those of the latter. Hence, the stability of the caged Sin clusters encapsulating Ho is related to the number of Ho-coordinated silicon atoms.
HoSi18. In this size, five stable cage-like HoSi18 isomers, 18a, 18b, 18c, 18d, and 18e, are obtained. The 18a isomer, the lowest-energy isomer and the most stable structure, is formed with one Si atom being surface-capped onto the deformed HoSi17 17b isomer whereas the 18b isomer is described as one Si atom being surface-capped onto the 17c isomer. The 18b isomer is higher in total energy by 0.553 eV than the 18a isomer. The 18c isomer is generated by capping one silicon atom on the surface of 17c while the 18d isomer can be seen as a rhombus being surface-capped on the deformed sandwich-like heptagonal prism D7h 14c. Furthermore, the sandwich-like geometry is higher in energy than the two connected Si10 units’ (16b-like) geometry. The 18e geometry deforms 18a geometry with the encapsulated Ho atom being localized at a different position; in other words, the encapsulated Ho atom in the 18a and 18e isomers interacts with different silicon atoms and has different coordination numbers; thus, the encapsulated Ho atom tries to terminate dangling bonds of different silicon atoms.
HoSi19. As far as the HoSi19 19a isomer is concerned, the geometry can be described as the Ho atom being encapsulated into the slightly distorted cage-like Si19 frame. The formed 19a isomer is composed of the two Si10 units and it is generated after a silicon atom is surface-added onto the 18a isomer. From the calculated total energies listed in Table 1, it is obvious that the 19a geometry is the lowest-energy isomer. The distorted C2v HoSi19 19b isomer is obtained after the Si atoms are surface-capped onto the distorted sandwich-like D6h HoSi12 frame. However, the 19a isomer is more stable than the 19b isomer. Guided by the above investigations of the 18d isomer, the striking 19c and 19d isomers are also considered. The C2v 19c isomer is depicted as the Ho atom being inserted into the slightly deformed heptagonal prism Si14 frame, together with a quadrangular pyramidal Si5 unit being capped on the side of the Si14 frame. It should be mentioned that the 19c isomer is higher in total energy than the 19a and 19b isomers by 2.65 and 0.11 eV, respectively. As a consequence, the layered 19c structure is not the most stable building block for the formation of the new material. The 19d isomer is depicted as the Ho atom being inserted into the layered sandwich-like octagonal prism HoSi16 unit. Interestingly, this structure is lower in stability than the 19a isomer and is not selected as the ground state and building block of new materials. Therefore, the sandwich-like 19c and 19d isomers are not the dominant geometries which are examined by the calculated total energies.
HoSi20. Starting from various initial structures, the HoSi20 cluster is calculated. Three isomers are identified for the Ho-doped Si20 clusters, which are the distorted closed cage-like D2d 20a isomer with the enclosed Ho atom acting as the centre of a symmetrical axis, the layered heptagonal prism Si14 unit with the enclosed Ho atom being localized at the centre site and a book-like Si6 unit being surface-capped on the heptagonal prism, 20b, and the seriously distorted closed cage-like hexagonal prism D6h 20c isomer, respectively (Fig. 1).

Two remarkable fullerene-like silicon isomers encapsulating Ho, namely 20a and 20c, are formed. The 20a isomer is obtained from the 19a isomer. It has a perfect fullerene-like geometry with Ho being localized at the centre of the Si20 frame. The Ho atom in 20a tries to interact with more silicon atoms and terminate the dangling bonds of the silicon atoms. Hence, the perfect closed fullerene-like 20a structure has enhanced stability and is appropriate for the building block of optical and optoelectronic materials. Interestingly, the encapsulated Ho atom in the 20a isomer, which has geometry similar to that of the Pd2-doped Si20 isomer,25 behaves as an acceptor of charges which is due to the tendency of Ho(6s24f11) to attain a completely filled 5d10 and 4f14 configuration.29,30 The 20c isomer is yielded after two silicon atoms are symmetrically capped on two sides of the 18b cluster. According to the calculated total energies listed in Table 1, it is obvious that 20c is higher in total energy than 20a by 3.95 eV. In addition, the 20b isomer is generated after a silicon atom is added on to the surface of the 19c isomer; moreover, the 20b isomer is less stable than the closed cage-like 20a cluster with a total energy difference of 3.34 eV.

3.2. Relative stabilities

In order to investigate the relative stabilities of the most stable HoSin (n = 12–20) clusters, the atomic averaged binding energies for the HoSin clusters (Eb(n)) and the fragmentation energies (D(n,n − 1)) with respect to removal of one Si atom from the most stable HoSin clusters are investigated (Fig. 2 and Table 2). The atomic averaged binding energies and fragmentation energies of the HoSin (n = 12–20) clusters are defined as:
Eb(n) = [ET(Ho) + nET(Si) − ET(HoSin)]/(n + 1)

D(n,n − 1) = ET(HoSin−1) + ET(Si) − ET(HoSin)
where ET(HoSin), ET(HoSin−1), ET(Si), and ET(Ho) denote the total energies of HoSin, HoSin−1, Si, and Ho clusters, respectively. The calculated results of the most stable HoSin (n = 12–20) isomers are plotted as the curves of Eb and D(n,n − 1) against the corresponding number of the Sin atoms (Fig. 2 and Table 2). It should be mentioned that the features of the size-evolution are intuitively viewed and the peaks of the curves correspond to those clusters having enhanced local stabilities.

image file: c4ra11828f-f2.tif
Fig. 2 The calculated averaged atomic binding energies (a) and fragmentation energies (b) for the most stable HoSin (n = 12–20) clusters.
Table 2 The calculated fragmentation energies and the atomic averaged binding energies of the most stable HoSin (n = 12–20) clusters (unit: eV)
System 12 13 14 15 16 17 18 19 20
Eb 2.78 2.80 2.80 2.85 2.90 2.88 2.95 2.94 3.01
D(n,n − 1)   3.12 2.81 3.53 3.74 2.55 4.17 2.87 4.36


The atomic averaged binding energies of the HoSin clusters are calculated and are listed in Table 2 and displayed in Fig. 2a; the curve generally shows the tendency for increased binding energy as the size of n is increased. As seen from Fig. 2a, it is noticed that three peaks with n = 16, 18, and 20 are found, reflecting that the predicted relative stabilities of the HoSin (n = 16, 18, and 20) clusters have stronger local stabilities as compared with the others.

According to the calculated fragmentation energies shown in Fig. 2b and Table 2, three remarkable peaks at n = 16, 18, and 20 for the HoSin (n = 12–20) clusters are found, showing that the corresponding clusters have stronger relative stabilities and have large abundances in mass spectroscopy as compared to the others. Moreover, the order of stabilities is HoSi20 > HoSi18 > HoSi16; this finding is in good agreement with the calculated atomic averaged binding energy. Interestingly, the Ho-doped silicon clusters have different magic numbers of stabilities as compared with the single transition metal doped silicon clusters.30 The 20a isomer in terms of the calculated fragmentation energies and averaged atomic binding energies shows the strongest geometrical stability, reflecting that the closed fullerene-like 20a geometry is the preferred structure. According to the calculated total energies of the HoSi20 20a and HoSi16 16a isomers, it is noticed that the 20a and 16a isomers can act as the building blocks for nanomaterials.

3.3. Charge-transfer mechanism

The calculated atomic polar tensor-based charges (APT) are tabulated in Table 1 and are displayed in Fig. 3. According to the calculated APT data for the Ho atom in the HoSin (n = 12–20) clusters, all the APT values of Ho are negative except for the APT value of the Ho atom in the 13a isomer (Table 1), showing that the electronic charges in the corresponding clusters transfer from the Sin frames to the Ho atom; and the unsaturated d orbitals of the Ho atom obtain the electrons from the Si atoms because the f orbitals are not involved in chemical bonds. These results coincide with the previous findings for ReSin and WSi15.13,15 As is the general trend, the APT populations of Ho are first decreased to a certain value and then increase. The two peaks at n = 16 and 19 correspond to the clusters with large charge-transfers and the charge-transfer of the HoSi16 cluster is the biggest one. It should be pointed out that the encapsulated Ho atom in the most stable HoSin (n = 12–20) clusters obtains more charges from its surroundings than the surface-capped Ho atom, and that the encapsulated Ho atom has a tendency to interact with more silicon atoms with inequivalent bond lengths and tries to terminate the dangling bonds of the silicon atoms. Therefore, the doped Ho atom plays very important roles in determining the stabilities of the Ho doped Sin (n = 12–20) clusters.
image file: c4ra11828f-f3.tif
Fig. 3 Size dependence of the APT of the most stable HoSin (n = 12–20) clusters.

3.4. HOMO and LUMO properties

The HOMO–LUMO gaps of the HoSin clusters are tabulated in Table 1. Based on our calculated results, the HOMO–LUMO gaps for the most stable HoSin (n = 12–17) clusters are about 1.5 eV while the HoSi16 cluster (1.95 eV) has the biggest HOMO–LUMO gap, indicating that the most stable HoSin (n = 12–17) clusters would be effective semiconductor materials (Fig. 4). In other words, the HoSi16 cluster is the strongest one, in terms of chemical stability, of all the HoSin (n = 12–17) clusters. Therefore, having a stronger relative stability and chemical stability, the HoSi16 cluster can be seen as the most suitable building block and can be selected as a candidate for novel nanomaterials. As far as the HoSi20 cluster is concerned, the HOMO–LUMO gap is 1.45 eV which is smaller than that of the HoSi16 (1.95 eV) cluster and therefore the HoSi16 cluster is stronger in terms of chemical stability than the HoSi20 cluster. Furthermore, we can predict that the HOMO–LUMO gaps for the HoSi20 and HoSi16 clusters can be detected experimentally. Consequently, the HoSi16 and HoSi20 clusters are stable in ionization and dissociation processes and contribute to forming the most preferred Ho-doped silicon nanoclusters, and make them more attractive for the cluster-assembled rare earth doped semiconductive nanomaterials with novel optical and optoelectronic properties.
image file: c4ra11828f-f4.tif
Fig. 4 Size dependence of the HOMO–LUMO gap of the most stable HoSin (n = 12–20) clusters.

The contour map of the HOMO for the most stable isomers (HoSi16, HoSi18, and HoSi20) is displayed in Fig. 5. It can be seen from Fig. 5 that some π-type and σ-type bonds are formed among the Sin atoms. The encapsulated Ho atom in the HoSi16 isomer interacts with 16 silicon atoms simultaneously and terminates the dangling bonds of the silicon atoms. As far as the HOMO property of the HoSi16 cluster is concerned, the π-type bonds are formed among the Sin atoms, which are distributed beside the Ho atom, and correspond to the localized π-type bonds. Furthermore, the encapsulated Ho atom obtains more charges from its silicon atoms and interacts with silicon atoms by relatively d–s σ-type bonds. From the contour maps of the HOMOs for the HoSi16 HoSi18, and HoSi20 clusters, it is obvious that the encapsulated Ho atom is surrounded by silicon atoms and only a few Si–Si bonds contribute to the π-type bonds because the σ-type bonds are the dominant Si–Si bonds in HoSi16 HoSi18, and HoSi20 clusters.


image file: c4ra11828f-f5.tif
Fig. 5 The contour maps of the HOMO for the most stable 16a, 18a, and 20a isomers.

3.5. Dipole moments

The dipole moments, which can be used to reflect the global geometry of the HoSin (n = 12–17) clusters, are calculated. The calculated results for the most stable Ho doped Sin clusters are tabulated in Table 1 and shown in Fig. 6. As seen from Fig. 6, it is apparent that the curve shows an oscillating tendency as the size of the silicon cluster increases. The three local minima are for HoSi16, HoSi18, and HoSi20 clusters, corresponding to dipole moments of 0.03, 0.31, and 0.0 Debye, respectively. HoSi16 and HoSi20 clusters are completely global geometries.
image file: c4ra11828f-f6.tif
Fig. 6 The calculated dipole moments of the most stable HoSin (n = 12–20) clusters.

4. Conclusions

The HoSin (n = 12–20) clusters with different spin states have been systematically investigated at the level of the DFT with the hybrid exchange and correlation (X3LYP) functional in combination with the 6-31G basis sets for the Si atoms and the Large-core Stuttgart effective core potentials (ECP56MHF) for the Ho element. The total energies, growth-pattern, equilibrium geometries as well as the APT charges of the HoSin (n = 12–20) clusters are calculated.

(1) The relative stabilities in terms of the calculated atomic averaged binding energies and fragmentation energies are discussed in detail. The calculated results reveal that the cake-like HoSin (n = 16, 18, 20) clusters have enhanced stabilities as compared with the others and HoSi20 is the most stable one. The order of stability is predicted to be HoSi16 < HoSi18 < HoSi20. Interestingly, the calculated dipole moments of the cake-like HoSin (n = 16, 18, 20) clusters are a lot smaller than that of the others, which correspond to the global minima; furthermore, the dipole moments for HoSi16 and HoSi20 are 0.03 and 0.00 Debye, respectively.

(2) The highest occupied molecular orbital – lowest unoccupied molecular orbital (HOMO–LUMO) gaps of the HoSin (n = 12–17) clusters are calculated. The calculated values for the HoSin (n = 12–17) clusters are above 1.55 eV while HoSi16 has the largest value (1.95 eV), reflecting that the HoSin (n = 12–17) cluster is a good semiconductor material, especially the HoSi16 cluster.

(3) According to the calculated APT charges of the Ho atom in the HoSin (n = 12–20) clusters, the contribution of charge-transfer to the stability of the HoSin clusters is briefly discussed, manifesting that the charges in the HoSin clusters transfer from the Si atoms to the Ho atom, and the charge-transfer in the HoSi16 cluster is the biggest one and has enhanced stability.

(4) The Ho atom in the lowest-energy HoSin (n = 12–14) isomers occupies the surface site while the optimized geometries show that the rare earth Ho atom in the HoSin (n ≥ 15) clusters is completely encapsulated into the Si frame. Furthermore, the critical size of the clusters required for the enclosed Ho atom to completely fall into the Sin (n = 12–20) frames is predicted to be n = 15; this finding is in good agreement with experimental results.5

Acknowledgements

This work is supported by the Natural Science fund of China (11179035, 10979048, 10875126), the 973 fund of the Chinese Ministry of Science and Technology (2010CB934504), the Innovation Program of Shanghai Municipal Education Commission (14YZ164 and 12YZ185) as well as Physical electronics disciplines (no. 12XKJC01).

References

  1. S. M. Beck, J. Chem. Phys., 1989, 90, 6306 CrossRef CAS PubMed.
  2. M. Ohara, K. Miyajima, A. Pramann, A. Nakajima and K. Kaya, J. Phys. Chem. A, 2002, 106, 3702 CrossRef CAS.
  3. K. Koyasu, J. Atobe, S. Furuse and A. I. Nakajima, J. Chem. Phys., 2008, 129, 214301 CrossRef PubMed.
  4. R. N. Zhao, Z. Ren, J. T. Bai, P. Guo and J. G. Han, J. Phys. Chem. A, 2006, 110, 4071 CrossRef CAS PubMed.
  5. A. Grubisic, Y. J. Ko, H. Wang and K. H. Bowen, J. Am. Chem. Soc., 2009, 131, 10783 CrossRef CAS PubMed.
  6. H. Q. Wang and H. F. Li, RSC Adv., 2014, 4, 56 Search PubMed.
  7. 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.
  8. R. N. Zhao, J. G. Han, J. T. Bai, F. Y. Liu and L. S. Sheng, Chem. Phys., 2010, 372, 89 CrossRef CAS PubMed.
  9. R. N. Zhao, J. G. Han and L. S. Sheng, Chem. Phys., 2010, 378, 82 CrossRef CAS PubMed.
  10. Z. L. Liu, H. K. Yuan and H. Chen, Appl. Mech. Mater., 2012, 130, 1035 Search PubMed.
  11. C. Xiao, A. Abraham and R. Quinn, et al., J. Phys. Chem. A, 2002, 106, 11380 CrossRef CAS.
  12. J. G. Han, C. Xiao and F. Hagelberg, Struct. Chem., 2002, 13, 173 CrossRef CAS.
  13. J. G. Han and Y. Y. Shi, Chem. Phys., 2001, 266, 33 CrossRef CAS.
  14. J. Wang and J. G. Han, J. Chem. Phys., 2005, 123, 244303 CrossRef PubMed.
  15. J. G. Han, Z. Y. Ren and B. Z. Lu, J. Phys. Chem. A, 2004, 108, 5100 CrossRef CAS.
  16. R. N. Zhao, Y. H. Yuan and J. G. Han, RSC Adv. Search PubMed submitted.
  17. J. Wang and J. G. Han, J. Phys. Chem. A, 2006, 110, 12670 CrossRef CAS PubMed.
  18. J. G. Han, Chem. Phys., 2001, 263, 255 CrossRef CAS.
  19. M. Dolg, H. Stoll, A. Savin and H. Preuss, Theor. Chim. Acta, 1989, 75, 173 CrossRef CAS.
  20. R. N. Zhao and Y. H. Yuan, J. Cluster Sci. Search PubMed , in press.
  21. R. N. Zhao, Y. H. Yuan, J. G. Han and Y. H. Duan, RSC Adv., 2014, 4, 59331 RSC.
  22. 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, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian09, Gaussian, Inc., Wallingford CT, 2009 Search PubMed.
  23. K. P. Huber and G. Herzberg, Constants of Diatomic Molecules, Van Nostrand Reinhold, New York, 1979 Search PubMed.
  24. J. G. Han, R. N. Zhao and Y. H. Duan, J. Phys. Chem. A, 2007, 111, 2148 CrossRef CAS PubMed.
  25. R. N. Zhao, J. G. Han and Y. H. Duan, Thin Solid Films, 2014, 556, 571 CrossRef CAS PubMed.
  26. G. Z. Zhang, MS thesis, Theoretical study on the interactions between small molecules and rare earth elements, Fudan, China, 2007, 24 Search PubMed.
  27. C. J. Barden, J. C. Rienstra-Kiracofe and H. F. Schaefer III, J. Chem. Phys., 2000, 113, 690 CrossRef CAS PubMed.
  28. J. G. Han and F. Hagelberg, J. Mol. Struct.: THEOCHEM, 2001, 549, 165 CrossRef CAS.
  29. J. G. Han, Chem. Phys., 2003, 286, 181 CrossRef CAS.
  30. J. G. Han and F. Hagelberg, J. Comput. Theor. Nanosci., 2009, 6, 257 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra11828f

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