A comparative ab initio study to investigate the rich structural variety and electronic properties of GamTen (m = 1, 2 and n = 1–4) with analogous oxides, sulfides and selenides

N. Seeburruna, I. A. Alswaidanb, H.-K. Funbc, E. F. Archibongd and P. Ramasami*a
aComputational Chemistry Group, Department of Chemistry, Faculty of Science, University of Mauritius, Réduit, Mauritius. E-mail: p.ramasami@uom.ac.mu; Fax: +230 4656928; Tel: +230 4657507
bDepartment of Pharmaceutical Chemistry, College of Pharmacy, King Saud University, P.O. Box 2457, Riyadh 11451, Saudi Arabia
cX-ray Crystallography Unit, School of Physics, Universiti Sains Malaysia, 11800 USM, Penang, Malaysia
dDepartment of Chemistry and Biochemsistry, University of Namibia, Namibia

Received 26th April 2015 , Accepted 31st July 2015

First published on 31st July 2015


Abstract

A series of gallium telluride, GamTen (m = 1, 2 and n = 1–4), clusters has been examined using density functional theory (DFT), second-order Møller–Plesset perturbation theory (MP2) and the coupled cluster approach with single and double substitutions and a perturbative treatment of triple excitations [CCSD(T)]. The study unravels the question of whether neutral GaTe2 is isostructural with BO2 as proposed earlier by a previous experiment (Trans. Faraday Soc. 1968, 64, 2998). The results of gallium tellurides are compared with the oxygen, sulfur and selenium analogues. In most cases, the substitution of O/S/Se by Te atoms in the gallium clusters does not drastically affect the structural characteristics. The adiabatic electron affinities (AEAs) of GamTen (m = 1, 2 and n = 1–4) clusters range from 1.33 to 3.46 eV at the CCSD(T)//B3LYP level. The AEAs of gallium tellurides are found to be independent on the electrophilicity of the clusters. Further, the adiabatic ionization potentials (AIPs) of the clusters are in good agreement with available experimental data. This research is expected to provide insight into the structural characteristics and electronic properties of gallium chalcogenides.


1. Introduction

Gallium chalcogenides have attracted scientific and commercial interest because they are semiconducting materials with extensive technological applications. The interaction between gallium and tellurium atoms is of prime importance due to their wide usage in optoelectronic and photovoltaic devices.1,2 Gallium monotelluride, GaTe is a heteronuclear gallium telluride which has been intensively studied.3 Owing to its band gap of around 1.7 eV, GaTe is useful in photoelectronic devices, radiation detectors,1,2 field-effect transistors and phototransistors.4 Doped gallium telluride clusters such as AgGaTe2 is used in thin film solar cells,5 CuGaTe2 as a high-performance thermoelectric material6 and LiGaTe2 for optical applications.7 Neutral telluroether complexes containing gallium atoms are used in chemical vapor deposition8,9 and as phase change memory materials.10

Extensive efforts have been devoted to the study of gallium chalcogenide clusters.1–24 Experiments and high level computations have shown GaO2 to be linear.22,23 However, theoretical results reveal that GaS2 and GaSe2 adopt cyclic geometries while anionic GaX2 (X = O–Se) prefers a linear configuration (D∞h).11–15 The lowest-energy structure of neutral and anionic GaX4 (X = O–Se) possesses a kite-shape with a trichalcogenide unit.14–16 Ga2Se and Ga2Te adopt bent configurations11,16 unlike Ga2O which is linear.12 For neutral and anionic Ga2X3 (X = O–Te),13,17,19,20 a V-shape and kite-shape are obtained, respectively. In the case of the Ga2X3 (X = O–Te)13,17,19,20 series, a decrease in the adiabatic electron affinity (AEA) values is noted at the CCSD(T)//B3LYP level of theory. Ga2X4 and their respective anions (X = O–Se) have similar ground state geometries.15,16,21 Structural differences are observed when Ga3X2 (X = O–Te) clusters are compared. Neutral Ga3O2 17 and Ga3S2 15 adopt kite-shape geometries while Ga3Se2 and Ga3Te2 adopt three-dimensional structures with D2h and C2v symmetries.20,22 In the same vein, Ga3S2, Ga3Se2 and Ga3Te216,19,24 prefer three-dimensional structures with a D3h symmetry differing from the planar kite geometry of Ga3O2.17

In view of the above, structural and electronic changes are highlighted by substitution of atoms. Fueled by these interesting results and an urge to see whether significant differences are noted on going down the periodic table, a systematic theoretical investigation of gallium tellurides, GamTen (m = 1, 2 and n = 1–4) was embarked to assess the structural and electronic properties. The objectives of the research are: (i) to study the equilibrium structures; (ii) to provide a reliable theoretical prediction of the relative stabilities, harmonic vibrational frequencies and energetic features such as electron detachment energies, electron affinities and ionization potentials (iii) to compare the ground state geometries and variations in the electronic properties with analogous oxides, sulfides and selenides and (iv) to compare the electron affinities of gallium telluride clusters with chlorine atom in order to determine whether they can be classified as superhalogens.

2. Computational methods

The computations of mono and digallium telluride, GamTen (m = 1, 2 and n = 1–4), clusters were performed with the Gaussian 0925 program. To obtain the ground state structures, various initial geometries were taken from isoelectronic gallium oxides, sulfides and selenides14–21 [Fig. 1–5]. Based on previous investigations,12,14,15–17,19–21 the 6-311+G(2df) one-particle basis set was employed for Ga atom while the LANL2DZdp ECP26 and def2-TZVP triple split valence basis sets27 were used for Te atom. The density functional theory (DFT) with the B3P86, B3PW91 and B3LYP functionals28–33 and the MP234,35 level of theory were employed. The single point CCSD(T)36 computations were performed using optimized geometries from the B3LYP functional.
image file: c5ra07594g-f1.tif
Fig. 1 Geometrical features of diatomic GaTe with the electronic states [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]. aB3P86, bB3PW91, cB3LYP and dMP2.

The adiabatic electron affinity (AEA), adiabatic detachment energy (ADE), vertical detachment energy (VDE), adiabatic ionization potential (AIP), vertical ionization potential (VIP) and chemical hardness (η) were calculated as follows:

AEA = E (optimized neutral at ground state) − E (optimized anion at ground state);

ADE = E (optimized corresponding neutral) − E (optimized anion);

VDE = E (neutral at optimized anion geometry) − E (optimized anion);

AIP = E (optimized cation at ground state) − E (optimized neutral at ground state);

VIP = E (cation at optimized neutral geometry) − E (optimized cation);

η ≈ (VIP − VDE)/2

The harmonic vibrational frequencies of the optimized geometries were analyzed to ensure that all the ground state structures belong to the minima on the potential energy surface. The Natural Bond orbital (NBO) analysis37,38 was performed with the B3LYP functional to provide insight into the nature of bonding in the clusters. The HOMO–LUMO gaps and chemical hardness (η) of the neutral gallium tellurides were also calculated with the B3LYP functional. In addition, the dissociation energies (De) of the studied clusters were carried out.

3. Results and discussion

Optimized geometrical configurations of the lowest-energy states of gallium telluride clusters, GamTen (m = 1, 2 and n = 1–4) are shown from Fig. 1–5. The energetic data (ΔE) is listed in Table 1. The internal coordinates and harmonic vibrational frequencies of the ground state geometries are available in Tables S1-S8 (ESI). The AEA and AIP values are presented in Tables 2 and 3. The dissociation energies (De) of the gallium tellurides are shown in Table 4. The natural bond orbital (NBO)34,35 on the gallium atoms are summarized in Table S9 while the VDE and chemical hardness (η) are given in Table S10.
Table 1 Energy shifts (eV) of the first low-lying states with respect to ground states of the mono and digallium tellurides [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]
Methods B3P86 B3PW91 B3LYP MP2
GaTe 0.08 0.08 0.01 0.20
GaTe 2.19 2.14 2.20 2.21
GaTe+ 0.94 0.97 0.99 0.79
GaTe2 0.18 0.18 0.18 0.01
GaTe2 0.93 0.92 0.92 1.33
GaTe2+ 0.35 0.35 0.34 0.35
GaTe3 0.12 0.12 0.19 0.60
GaTe3 0.80 0.81 0.71 1.45
GaTe3+ 0.14 0.15 0.05 0.83
GaTe4 0.09 0.07 0.17 −0.23
GaTe4 0.21 0.19 0.28 −0.02
GaTe4+ 0.20 0.21 0.11 0.63
Ga2Te 0.14 0.14 0.13 0.19
Ga2Te 0.49 0.48 0.42 0.45
Ga2Te+ 0.12 0.13 0.12 0.28
Ga2Te2 0.57 0.57 0.49 0.33
Ga2Te2 0.62 0.61 0.54 0.65
Ga2Te2+ 0.39 0.39 0.38 0.43
Ga2Te3+ 0.19 0.20 0.17 0.67
Ga2Te4 0.17 0.18 0.28 0.44
Ga2Te4 0.23 0.23 0.22 0.60
Ga2Te4+ 0.19 0.20 0.12 0.77


Table 2 Adiabatic electron affinities (AEAs) of the gallium telluride clusters using different levels of theory [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]
Cluster B3P86 B3PW91 B3LYP MP2 CCSD(T)//B3LYP
a AEA values in parenthesis were obtained using single point CCSD(T) with def2-TZVP basis set.b AEA value in bold was obtained at the CCSD(T) level of theory.
GaTe 3.20 2.65 2.67 2.48 2.46 (2.57)a, 2.46b
GaTe2 3.83 3.29 3.23 3.55 3.46 (3.38)
GaTe3 3.94 3.40 3.31 3.20 3.23 (3.50)
GaTe4 3.81 3.27 3.24 3.23 3.10 (3.30)
Ga2Te 1.92 1.39 1.25 1.32 1.33 (1.37)
Ga2Te2 2.95 2.40 2.35 2.49 2.46 (2.34)
Ga2Te3 3.71 3.18 3.02 2.77 2.77 (3.10)
Ga2Te4 4.07 3.51 3.50 3.14 3.01 (3.41)


Table 3 Adiabatic ionization potentials (AIPs) of the gallium telluride clusters using different levels of theory [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]
Cluster B3P86 B3PW91 B3LYP MP2 CCSD(T)//B3LYP
a AIP values in parenthesis were obtained using single point CCSD(T) with def2-TZVP basis set.b AIP value in bold was obtained at the CCSD(T) level of theory.
GaTe 8.82 8.25 8.17 8.08 7.85 (8.20)a, 8.01b
GaTe2 8.36 7.80 7.65 7.73 7.84 (7.75)
GaTe3 7.84 7.29 7.36 7.08 6.85 (7.13)
GaTe4 7.64 7.08 7.14 6.45 6.52 (6.92)
Ga2Te 8.21 7.65 7.62 7.54 7.55 (7.63)
Ga2Te2 8.22 7.65 7.51 7.27 7.13 (7.64)
Ga2Te3 7.66 7.10 7.22 6.92 6.86 (7.03)
Ga2Te4 7.54 7.00 7.10 6.80 6.93 (6.95)


Table 4 Dissociation energies (De, kJ mol−1) of gallium tellurides through different channels [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]
Channels B3P86 B3PW91 B3LYP MP2 CCSD(T)//B3LYP
GaTe → Ga + Te 282.3 272.6 259.6 266.2 260.5
GaTe2 → Ga + Te2 266.5 258.4 243.8 275.7 265.9
GaTe2 → GaTe + Te 274.7 266.2 252.9 234.9 224.1
GaTe2 → Ga + 2Te 557.0 538.8 512.5 501.0 484.7
GaTe3 → GaTe + Te2 169.5 163.4 144.2 217.3 192.1
GaTe3 → GaTe2 + Te 185.4 177.6 159.9 207.7 186.7
GaTe3 → GaTe + 2Te 460.1 443.8 412.8 442.6 410.8
GaTe4 → GaTe2 + Te2 139.2 131.9 105.9 196.9 171.4
GaTe4 → GaTe + Te3 232.4 225.8 207.7 294.2 264.8
GaTe4 → GaTe3 + Te 244.3 234.6 214.6 214.2 203.4
GaTe4 → Ga + 2Te2 405.7 390.3 349.7 472.6 437.3
Ga2Te → GaTe + Ga 293.1 285.4 282.6 284.7 276.3
Ga2Te → 2Ga + Te 575.4 558.0 542.2 550.9 536.8
Ga2Te2 →2GaTe 265.1 258.8 242.2 236.6 235.9
Ga2Te2 → Ga2Te + Te 254.4 246.0 219.1 238.1 220.2
Ga2Te2 → GaTe2 + Ga 272.8 262.2 248.9 288.0 272.3
Ga2Te3 → Ga2Te + Te2 172.6 166.3 136.0 278.6 253.6
Ga2Te3 → GaTe + GaTe2 199.1 193.3 174.8 287.6 263.9
Ga2Te3 → Ga2Te2 + Te 208.7 200.7 185.5 265.8 252.2
Ga2Te3 → Ga2Te + 2Te 463.1 446.7 404.6 503.9 472.3
Ga2Te4 → Ga2Te2 + Te2 165.5 161.2 131.1 279.1 272.4
Ga2Te4 → 2GaTe2 171.2 168.1 136.1 291.3 278.8
Ga2Te4 → GaTe3 + GaTe 261.1 256.6 229.1 318.4 316.2
Ga2Te4 → Ga2Te3 + Te 247.3 241.0 214.3 238.6 239.0
Ga2Te4 → Ga2Te2 + 2Te 456.0 441.6 399.7 504.3 491.1


A. Structural properties

The electronic ground state of neutral GaTe is 2Σ+, which is in agreement with GaO39 and GaSe.16 The quartet and sextet states of GaTe are well separated from the doublet ground state by 2.68 and 5.23 eV, respectively at the CCSD(T)//B3LYP level. The most stable electronic states of GaTe and GaTe+ are 1Σ+ and 3Σ+ (Fig. 1), respectively. This result is consistent with anionic and cationic GaO39 and GaSe.16 A cyclic structure is found as the lowest-energy structure for GaTe2 with the DFT and MP2 methods. However, single point CCSD(T) using DFT and MP2 geometries, reveals a linear structure (Fig. 2) by analogy with BO2.11 This ambiguity was solved with def2-TZVP,27 being a large basis set. Single point CCSD(T) computations with def2-TZVP basis set reveals a cyclic structure for GaTe2. Furthermore, the linear structure is a saddle point with the MP2 method. In this vein, the linear structure of GaTe2 is the first low-lying isomer. For the tri-atomic series, GaTe2 is cyclic as GaS2 15 and GaSe2 16 while GaO2 is linear.22,23
image file: c5ra07594g-f2.tif
Fig. 2 Geometrical features of GaTe2 and Ga2Te with the electronic states [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]. aB3P86, bB3PW91, cB3LYP and dMP2.

Akin to GaO2,12 GaS2[thin space (1/6-em)]15 and GaSe2,16 anionic GaTe2 prefers a linear centrosymmetric (D∞h) configuration. The first low-lying geometry of GaTe2 is bent (1A1). As its selenide congener,16 cationic GaTe2 adopts a bent configuration with a Cs symmetry. The lowest-energy configuration of neutral, negatively and positively charged GaTe3 consists of a diatomic GaTe molecule bound perpendicularly to a Te2 moiety (Fig. 3). Similar neutral and anionic ground state geometries were reported for gallium selenide clusters.16 Neutral, negatively and positively charged GaTe3 adopt a rhombic structure as the first low lying isomer. GaTe4 and its anion adopt kite geometries with a tritelluride unit (2B1) analogous to gallium oxide,14 sulfide15 and selenide.16 This result is further confirmed by the CCSD(T)//B3LYP and CCSD(T)//MP2 computations. The low-lying structure of neutral and anionic GaTe4 is of D2d symmetry, which is the lowest-energy structure for GaTe4+. On the other hand, the low-lying structure of GaTe4+ adopts the kite geometry which is 0.63 eV above the ground state geometry with the MP2 level.


image file: c5ra07594g-f3.tif
Fig. 3 Geometrical features of GaTe3 and Ga2Te2 with the electronic states [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]. aB3P86, bB3PW91, cB3LYP and dMP2.

The lowest-energy state of neutral Ga2Te is 2A2 with a bent configuration (C2v) (Fig. 2). This result is consistent with Uy et al.11 This geometry is isostructural with Ga2Se16 but differs from Ga2O, which is linear.12 The next low-lying geometry of Ga2Te is linear. Anionic and cationic Ga2Te maintain the bent configuration of the neutral species but with C2v and Cs symmetries, respectively. Same behavior was pointed out with gallium selenide.16 A distinct difference is noticed within the Ga2X2 (X = O–Te) family as Ga2O2 is linear,12 Ga2S2 is rhombic,15 Ga2Se2 is L-shape16 and Ga2Te2 has a butterfly structure with a C2v symmetry (Fig. 3). The structure of Ga2Te2 is rhombic, which is analogous to Ga2O2,12 Ga2S2 15 and Ga2Se2.16 For Ga2Te2+, a L-shape structure is obtained as Ga2Se2+.16 The low-lying isomers of neutral, anionic and cationic Ga2Te2 adopt linear, cyclic and L-shape configurations, respectively. The lowest-energy structure of neutral and anionic Ga2Te4 can be viewed as rhombic with one terminal tellurium atom attached to each gallium atom (D2h) (Fig. 5). Similar ground state geometry was found with gallium oxide,21 sulfide15 and selenide.16 Ga2Te4+ adopts a similar ‘fish-like’ structure as Ga2Se4+.16 The next energetically low-lying isomer of Ga2Te4 is a ‘fish-like’ structure which is 0.44 eV above the ground state geometry (MP2). The low-lying isomer of anionic Ga2Te4 is a twisted hexagon configuration and cationic Ga2Te4 prefers a planar configuration with D2h symmetry. Gallium telluride clusters prefer planar structures with the exception of mono and digallium tetratelluride cations. A structural evolution is revealed upon addition of a tellurium atom to the GaTen and Ga2Ten series (Fig. S1 and S2). The NBO charges on gallium atoms are positive with the exception of GaTe and Ga2Te4+ (Table S9). No significant pattern is observed in the natural charge of gallium atom upon sequential addition of tellurium atoms with the exception of GaTen+ and Ga2Ten.


image file: c5ra07594g-f4.tif
Fig. 4 Geometrical features of GaTe4 and Ga2Te3 with the electronic states [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]. aB3P86, bB3PW91, cB3LYP and dMP2.

image file: c5ra07594g-f5.tif
Fig. 5 Geometrical features of Ga2Te4 with the electronic states [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp]. aB3P86, bB3PW91, cB3LYP and dMP2.

B. Geometrical properties

Akin to GaO39 and GaSe,16 a slight elongation of 0.072 Å (B3LYP) and 0.095 Å (MP2) is observed in the monotelluride Ga–Te bond length from GaTe to GaTe. As expected, an increase in the bond angle at the apex is observed with the cyclic structure of GaO2 (38.0°),12 GaS2 (47.5°),15 GaSe2 (51.0°)16 and GaTe2 (55.4°) [B3LYP]. Unlike the GaX3/GaX3 (X = O and Se)16,18 system, addition of an electron to GaTe3 shortens the terminal Ga–Te bond length (DFT and MP2). On the other hand, there is an increase in the terminal Ga–Te bond from GaTe3 to GaTe3+. Like gallium selenide,16 the bridge Ga–Te bond is elongated whereas the terminal Ga–Te bond is shortened from GaTe4 to GaTe4 (DFT and MP2). However, the reverse pattern was observed with the gallium oxide14 and sulfide15 congener. For the GaX4 (X = O–Te)14–16 series, a compression in the bond angle of the trichalcogenide unit is noted in all cases. Turning to the Ga2Ten series, the Ga–Te–Ga bond angle compresses (DFT and MP2) from Ga2Te to Ga2Te. Similar finding was found with gallium oxide12 and selenide.16 Upon addition of an electron to Ga2Te2, the three-dimensional structure turns into a planar configuration unlike Ga2Se2 16 where both neutral and cation maintain the planar L-shaped geometry. The bridge Ga–Te bonds of Ga2Te4 are slightly shorter than the anion whereas the terminal Ga–Te bonds are longer. The bond angle between the terminal Ga–Te and neighboring tellurium atom of Ga2Te4 is smaller than that of Ga2O4,21 Ga2S4 15 and Ga2Se4.16

The Ga–Te bond lengths of the neutral gallium tellurides are within the range 2.357–2.845 Å (B3P86), 2.331–2.851 Å (B3PW91), 2.342–2.884 Å (B3LYP) and 2.316–2.791 Å (MP2). The Ga–Te bond lengths of the gallium tellurides are in agreement with mixed chalcogenide cubanes,40 KGa2Te6,41 monoclinic GaTe,42,43 hexagonal GaTe,44 Ga2Te5 45 and telluroether gallium complexes.9 The Te–Te bond lengths of the neutral gallium tellurides are within the range 2.662–2.776 Å (B3P86), 2.663–2.777 Å (B3PW91), 2.680–2.827 Å (B3LYP) and 2.681–2.905 Å (MP2). The Te–Te bond lengths of the gallium tellurides are smaller than that of contact ion pair of tellurium system.45

C. Vibrational properties

All the ground state geometries of the gallium telluride clusters have real frequencies. The frequencies of GaTe with DFT functionals (257 cm−1 and 242 cm−1) agree very well with Uy et al. (250 cm−1).11 Like GaSe,16 the stretching frequency of GaTe and GaTe are close to each other. From Uy et al.11 investigation, GaTe2 has three vibrational frequencies 152, 104 and 92 cm−1, respectively. With linear GaTe2, four wavenumbers were achieved and this feature negates its possibility as ground state geometry. The theoretical vibrational frequencies are higher than the experimental values with the DFT and MP2 methods. The highest frequency mode of GaTe3 is attributed to the symmetrical stretching of terminal Ga–Te bond. Same observations were noted with GaO3[thin space (1/6-em)]18 and GaSe3.16 The highest frequency mode of GaX4 (X = O–Te)12,14–16 corresponds to the stretching of X–X bond. Turning to digallium chalcogenides, the most active mode of Ga2X (X = O, Se, Te)12,16 is the asymmetrical stretching of Ga–X bond. However, the harmonic vibrational frequencies of Ga2Te are not in good correlation with the values reported by Uy et al.11 Smaller wavenumbers are obtained with the DFT and MP2 levels of theory (Tables S2, S4, S6 and S8). The highest frequency vibrations of Ga2X2 and Ga2X4 (X = O–Te)12,15,16,21 correspond to the stretching of X–X and terminal Ga–X bond, respectively. No significant pattern is observed in the frequencies upon sequential addition of tellurium atom to the GaTen and Ga2Ten (n = 1–4) series.

D. Electronic properties

I. Adiabatic electron affinities (AEAs). The AEA of GaTe ranges from 2.46 eV [CCSD(T)//B3LYP] to 3.20 (B3P86). The AEA of GaTe at the CCSD(T) level is the same as reported at the single point CCSD(T)//B3LYP with LANL2DZdp basis set. The ground state electronic configuration of GaTe2 is (6σg)2(6πu)2(4πg)2(4σu)2(2δg)2. The AEA of GaTe2 varies from 3.23 eV (B3LYP) to 3.83 eV (B3P86). The excess electron of GaTe2 is distributed over the tellurium atoms (Fig. S3). Similar HOMO plots were observed for GaX2 (X = S and Se).15,16 The electronic configuration of GaTe3 is (12a1)2(6b2)2(5b1)2(2a2)2 and that of GaTe4 is (13a1)2(7b2)2(6b1)2(2a2)2. Their AEA values range from 3.20 to 3.94 and 3.10 to 3.81 eV, respectively. Akin to gallium selenide,16 the excess electron of GaTe3 and GaTe4 is localized on the terminal tellurium atom (Fig. S3).

The electronic configuration of Ga2Te is (14a1)1(12b2)2(5b1)2(4a2)2 and the electron detachment process of 1A1 (C2v) + e2A1 (C2v) involves the removal of an electron from the 14a1 molecular orbital (MO) to yield the 1A1 (C2v) ground state of Ga2Te. The AEAs of Ga2Te vary from 1.25 eV (B3LYP) to 1.92 eV (B3P86). The lowest-lying doublet state of Ga2Te2 has an electronic configuration of (10ag)2(9b2u)1(5b1u)2(4b3u)2(4b3g)2(3b1g)2(2b2g)2(1au)2 and the AEAs vary from 2.34 [CCSD(T)//B3LYP] to 2.95 eV (B3P86). Like the sulfur15 and selenium16 analogies, the extra electron of Ga2Te2 is distributed over the bridge tellurium atoms (Fig. S4). The electronic configuration of anionic Ga2Te4 is (11ag)2(10b1u)2(6b2u)2(5b3g)1(4b2g)2(2b3u)2(2b1g)2(1au)2 and transition 3B1u (D2h) + e2B2g (D2h) involves the removal of a b3g electron from the anion to produce the 3B1u (D2h) ground state of Ga2Te4. The AEAs range from 3.01 eV to 4.07 eV. For the Ga2X4 (X = O–Te)15,16,21 series, the extra electron of the anion is localized around the terminal chalcogen atoms (Fig. S4). The ADEs and AEAs are the same for the studied gallium telluride clusters, with the exception of Ga2Te2 and Ga2Te3, because of the similar ground state geometries of the neutral and anion.

A significant increase in the AEA values is seen when progressing from Ga2Te to Ga2Te4 (Fig. 6). This trend was observed with gallium oxides,21 sulfides15 and selenides.16 Similar increase in the AEAs is expected for the GaTen series but instead, a drop in the electron affinity values is obtained for GaTe3 and GaTe4 (Fig. 6). Among the gallium tellurides, GaTe2 has the highest electron affinity at the CCSD(T)//B3LYP level of theory. However, the electron affinity of GaTe2 is less than chlorine atom (3.62 eV)46 with different levels of theory and it has a VDE of 3.42 eV with the B3LYP functional. This is obvious due to electronegativity values. In this vein, unlike analogous gallium oxide,12 sulfide15 and selenide,16 GaTe2 cannot exhibit superhalogen properties. The AEA of the Ga2X2 (X = O–Te) system is the lowest among the series studied (Fig. S5). No significant trend is observed in the AEA values, with of exception Ga2X3, when changing the O/S/Se substituents to Te.12–17,19–22 This clearly demonstrates that the electron affinities of gallium chalcogenide clusters are independent of electronegative atoms in the system but are related to the electronic structures.


image file: c5ra07594g-f6.tif
Fig. 6 Electron affinities of the gallium telluride clusters at the CCSD(T)//B3LYP level [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp].
II. Adiabatic ionization potentials (AIPs). The calculated AIPs of GaTe and GaTe2 are in accordance with the available experimental data11. In the case of GaTe, the B3PW91 functional provides a reliable theoretical value while for GaTe2, the B3P86 yields a value closer to the experimental one. The AIP values of GaTe3 range from 6.85 [CCSD(T)//B3LYP] to 7.84 (B3P86) eV and GaTe4 from 6.45 (MP2) to 7.64 (B3P86) eV. Turning to digallium telluride series, the AIP of Ga2Te ranges from 7.54 (MP2) to 8.21 eV (B3P86), Ga2Te2 from 7.13 (CCSD(T)//B3LYP) to 8.22 (B3P86) eV, Ga2Te3 from 6.86 [CCSD(T)/B3LYP] to 7.66 (B3P86) eV and Ga2Te4 from 6.80 (MP2) to 7.54 (B3P86) eV. As expected, the AIP results of Ga2Te and Ga2Te2 are consistent with Uy et al.11 experimental data. The AIPs of the gallium tellurides are lower than the gallium selenides.16 As shown in Fig. 7, the AIPs decrease upon sequential addition of tellurium atom to the GaTen series. This shows that AIP is not only related to the total number of tellurium atoms but also on electronic structures. Similar observation was highlighted with AEA.
image file: c5ra07594g-f7.tif
Fig. 7 Ionization potentials of the gallium telluride clusters at the CCSD(T)//B3LYP level [basis sets Ga: 6-311+G(2df) and Te: LANL2DZdp].
III. Mean unsigned error (MUE). The accuracies of the different methods used in the research were examined by calculating the mean unsigned error (MUE).47 The single point CCSD(T)//B3LYP values with the def2-TZVP basis set were taken as reference data. With the AEA values, the MUE was 0.56 eV (B3P86), 0.07 eV (B3PW91), 0.10 eV (B3LYP), 0.18 eV (MP2) and 0.19 eV CCSD(T)//B3LYP with the LANL2DZdp basis set. Similarly, the MUE of the studied gallium tellurides was calculated by using the AIPs values. The calculated MUE is 0.85 eV (B3P86), 0.07 eV (B3PW91), 0.13 eV (B3LYP), 0.17 eV (MP2) and 0.23 eV CCSD(T)//B3LYP with the LANL2DZdp basis set. In both cases, the B3PW91 functional gives meaningful values.
IV. HOMO–LUMO gaps. The HOMO–LUMO gaps of neutral gallium telluride clusters are shown in Table S10. A large value of the HOMO–LUMO energy gap enhances chemical stability of the cluster.48 The HOMO–LUMO gaps for the lowest-energy configurations are large, varying from 2.61 (GaTe4) to 3.99 eV (GaTe) with the B3LYP functional. For the GaTen series, a general decrease in the HOMO–LUMO gaps is noted with an increase in the tellurium-to-metal ratio. However, no correlation is observed for the Ga2Ten series. Even though the substitution of O/S/Se by Te does not lead to a significant trend for the HOMO–LUMO energy gap, a decrease in the HOMO–LUMO energy gap values is noted for the Ga2X3 and Ga2X4[thin space (1/6-em)]15,16,18,21 series.
V. Chemical hardness (η). Pearson states: ‘Clusters arrange their electronic structures so as to have the maximum possible hardness’.49 The chemical hardness of the gallium tellurides varies from 1.58 to 3.19 eV with the B3LYP functional (Table S10). An increase in the tellurium-to-metal ratio decreases the chemical hardness of the Ga2Ten clusters. The HOMO–LUMO gaps can be related to hardness (η). Clusters with small HOMO–LUMO gaps are said to be ‘soft’.50 As mentioned earlier, the HOMO–LUMO gaps of some studied gallium tellurides are relatively large and thereby they can be considered as ‘hard’ clusters.

E. Thermodynamic stability

The dissociation energy (De) is obtained as the difference in total energies of the initial state and the sum of total energies of the decay fragments. The dissociation energies are positive indicating that the clusters are stable (Table 4). The De values of GaTe are in agreement with the experimental data (222–272 kJ mol−1).11 To compare the chalcogen effect for GaX (X = O–Te) series, the De of GaO and GaS were calculated. The De of GaX16 decreases with increasing atomic number of the chalcogen element. This implies that it is easier to cleave a Ga–Te than a Ga–O bond. The monogallium tellurides, with the exception of GaTe, favour a cascade release of tellurium molecule. The dissociation of Ga2Te4 results into Ga2Te2 and a Te molecule. This is because the terminal tellurium atoms are cleaved easier compared to bridge tellurium ones. Similar feature was observed for gallium sulfide and selenide.15,16,21 The De of the preferred channel decreases upon sequential addition of tellurium atom to GaTen and Ga2Ten series. Similar observation was seen with the selenium analogue.16

4. Conclusions

Using a number of computational methods, the structural and electronic properties of gallium telluride GamTen (m = 1, 2 and n = 1–4) clusters were investigated. One prime focus was to explore the substitution of O/S/Se by Te in the gallium clusters. The substitution of O/S/Se by Te in the studied gallium tellurides does not induce much structural changes. A structural divide is observed for the Ga2X2 (X = O–Te) series where Ga2O2 is linear, Ga2S2 is rhombic, Ga2Se2 is L-shape and Ga2Te2 has a butterfly structure. The reported Ga–Te bond lengths, AIPs and De are found to be in agreement with experimental data. The B3PW91 functional was found to provide more reliable data as it gives a smaller mean unsigned error. This is further confirmed from AIP values. Akin to gallium selenide clusters, the AEAs and AIPs of gallium tellurides are not solely related to the total number of tellurium atoms but also dependant on electronic structures. In contrast to analogous GaO2, GaS2 and GaSe2, GaTe2 is not classified as a superhalogen because it has lower AEA and VDE values than chlorine atom. The results of this research show that the substitution of an atom by another can open the door to novel structural and electronic properties.

Acknowledgements

N. S. acknowledges support from the Mauritius Tertiary Education Commission (TEC). The authors also acknowledge facilities from the University of Mauritius and the University of Namibia. The authors extend their appreciation to the Deanship of Scientific Research at King Saud University for the research group Project No. RGP VPP-207. The authors would like to thank the anonymous reviewers for useful comments to improve the manuscript.

References

  1. O. Madelung and R. Poerschke, Semiconductors: Other than Group IV Elements and III-V Compounds, Springer-Verlag, Berlin, 1992 Search PubMed.
  2. M. A. Malik, M. Afzaal and P. O' Brien, Chem. Rev., 2010, 110, 4417 CrossRef CAS PubMed.
  3. A. Zubiaga, F. Plazaola, J. A. García, C. Martínez-Tomás and V. Muñoz-Sanjosé, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 81, 195211 CrossRef.
  4. Z. Wang, K. Xu, Y. Li, X. Zhan, M. Safdar, Q. Wang, F. Wang and J. He, ACS Nano, 2014, 8, 4859 CrossRef CAS PubMed.
  5. K. Aravinth, A. G. Babu and P. Ramasamy, Mater. Sci. Semicond. Process., 2014, 24, 44 CrossRef CAS PubMed.
  6. K. Kurosaki and S. Yamanaka, Phys. Status Solidi A, 2013, 210, 82 CrossRef CAS PubMed.
  7. L. Isaenko, P. Krinitsin, V. Vedenyapin, A. Yelisseyey, A. Merkulov, J.-J. Zondy and V. Petrov, Cryst. Growth Des., 2005, 5, 1325 CAS.
  8. K. George, C. H. K. de Groot, C. Gurnani, A. L. Hector, R. Huang, M. Jura, W. Levason and G. Reid, Phys. Procedia, 2013, 46, 142 CrossRef CAS PubMed.
  9. K. George, C. H. K. de Groot, C. Gurnani, A. L. Hector, R. Huang, M. Jura, W. Levason and G. Reid, Chem. Mater., 2013, 25, 1829 CrossRef CAS.
  10. H. Zhu, J. Yin, Y. Xia and Z. Liu, Appl. Phys. Lett., 2010, 97, 083504 CrossRef PubMed.
  11. O. M. Uy, D. W. Muenow, P. J. Ficalora and J. L. Margrave, Trans. Faraday Soc., 1968, 64, 2998 RSC.
  12. S. Gowtham, A. Costales and R. Pandey, J. Phys. Chem. B, 2004, 108, 17295 CrossRef CAS.
  13. J. J. BelBruno, E. Sanville, A. Burnin, A. K. Muhangi and A. Malyutin, Chem. Phys. Lett., 2009, 478, 132 CrossRef CAS PubMed.
  14. E. F. Archibong and P. Ramasami, Comput. Theor. Chem., 2011, 964, 324 CrossRef CAS PubMed.
  15. N. Seeburrun, H. H. Abdallah, E. F. Archibong and P. Ramasami, Struct. Chem., 2014, 25, 755 CrossRef CAS.
  16. N. Seeburrun, E. F. Archibong and P. Ramasami, J. Mol. Model., 2015, 21, 42 CrossRef PubMed.
  17. E. F. Archibong and E. N. Mvula, Chem. Phys. Lett., 2005, 408, 371 CrossRef CAS PubMed.
  18. S. Gowtham, M. Deshpande, A. Costales and R. Pandey, J. Phys. Chem. B, 2005, 109, 14836 CrossRef CAS PubMed.
  19. N. Seeburrun, E. F. Archibong and P. Ramasami, Chem. Phys. Lett., 2008, 467, 23 CrossRef CAS PubMed.
  20. N. Seeburrun, H. H. Abdallah, E. F. Archibong and P. Ramasami, Eur. Phys. J. D, 2011, 63, 351 CrossRef CAS PubMed.
  21. N. Seeburrun, H. H. Abdallah and P. Ramasami, J. Phys. Chem. A, 2012, 116, 3215 CrossRef CAS PubMed.
  22. A. Köhn, B. Gaertner and H.-J. Himmel, Chem.–Eur. J., 2005, 11, 5575 CrossRef PubMed.
  23. B. Gaertner, A. Köhn and H.-J. Himmel, Eur. J. Inorg. Chem., 2006, 11, 1496 CrossRef PubMed.
  24. N. Seeburrun, M. M. J. Soopramanien, H. H. Abdallah, E. F. Archibong and P. Ramasami, J. Mater. Sci., 2012, 47, 4332 CrossRef CAS.
  25. 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 D.01, Gaussian, Inc., Wallingford CT, 2009 Search PubMed.
  26. S. Sriram and R. Chandiramouli, Res. Chem. Intermed., 2013, 41, 2095 CrossRef.
  27. K. Kunkel, E. Milke and M. Binnewies, Eur. J. Inorg. Chem., 2015, 1, 124 CrossRef PubMed.
  28. J. P. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 33, 8822 CrossRef.
  29. J. P. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 34, 7406 CrossRef.
  30. J. P. Perdew and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 45, 13244 CrossRef.
  31. 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.
  32. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785 CrossRef CAS.
  33. A. D. Becke, J. Chem. Phys., 1993, 98, 5648 CrossRef CAS PubMed.
  34. R. J. Bartlett, Annu. Rev. Phys. Chem., 1981, 32, 359 CrossRef CAS.
  35. W. J. Hehre, L. Radom, P. V. R. Schleyer and J. A. Pople, Ab Initio Molecular Orbital Theory, Wiley, New York, 1986 Search PubMed.
  36. K. Raghavachari, G. W. Trucks, J. A. Pople and M. Head-Gordon, Chem. Phys. Lett., 1989, 157, 479 CrossRef CAS.
  37. A. E. Reed, R. B. Weinstock and F. Weinhold, J. Chem. Phys., 1985, 83, 735 CrossRef CAS PubMed.
  38. A. E. Reed, L. A. Curtiss and F. Weinhold, Chem. Rev., 1988, 88, 899 CrossRef CAS.
  39. G. Meloni, S. M. Sheehan and D. M. Neumark, J. Chem. Phys., 2005, 122, 074317 CrossRef PubMed.
  40. B. D. Fahlman, A. Daniels, G. E. Scusceria and A. R. Barron, Dalton Trans., 2001, 3239 RSC.
  41. B. Eisenmann and A. Z. Hofmann, Kristallgor., 1991, 197, 145 CrossRef CAS.
  42. M. Julien-Pouzol, S. Jaulmes, M. Guittard and F. Alapini, Acta Crystallogr., 1979, 35, 2848 CrossRef.
  43. F. Alapini, J. Flahaut, M. Guittard, S. Jaulmes and M. Julien-Pouzol, J. Solid State Chem., 1979, 28, 309 CrossRef CAS.
  44. S. A. Semiletov and V. A. Vlasov, Kristallografiya, 1963, 8, 877 CAS.
  45. T. Chivers, J. S. Ritch, S. D. Robertson, J. Konu and H. M. Tuononen, Acc. Chem. Res., 2010, 43, 1053 CrossRef CAS PubMed.
  46. H. Hotop and W. C. Lineberger, J. Phys. Chem. Ref. Data, 1985, 14, 731 CrossRef CAS PubMed.
  47. Y. Zhao and D. G. Truhlar, J. Chem. Phys., 2006, 124, 224105 CrossRef PubMed.
  48. M. M. Zhong, X. Y. Kuang, Z. H. Wang, P. Shao and L. P. Ding, J. Mol. Model., 2013, 19, 263 CrossRef CAS PubMed.
  49. R. G. Pearson, Chemical hardness: Applications from molecules to solids, Wiley-VCH, Winheim, 1997 Search PubMed.
  50. R. G. Parr and P. K. Chattaraj, J. Am. Chem. Soc., 1991, 113, 1854 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Tables containing harmonic vibrational frequencies, ADEs, VDEs, VIPs (eV) and chemical hardness (η) of gallium telluride clusters. Figures presenting the structural evolution of mono and digallium tellurides. See DOI: 10.1039/c5ra07594g

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