Carbon monoxide interactions with pure and doped B11XN12 (X = Mg, Ge, Ga) nano-clusters: a theoretical study

Alireza Soltani*ab and Masoud Bezi Javanc
aJoints, Bones and Connective Tissue Research Center, Golestan University of Medical Science, Gorgan, Iran. E-mail: Alireza.soltani46@yahoo.com; Tel: +98-938-4544921
bYoung Researchers and Elite Club, Gorgan Branch, Islamic Azad University, Gorgan, Iran
cPhysics Department, Faculty of Sciences, Golestan University, Gorgan, Iran

Received 29th June 2015 , Accepted 8th September 2015

First published on 8th September 2015


Abstract

The goal of this investigation was to study a novel sensor for detecting the toxic gas compounds of CO using B11XN12 (X = Ge, Mg, and Ga) nano-clusters in terms of its energetic, geometric, and electronic structure using DFT calculations by the PBE-D method. The reaction of CO gas with these doping atoms results in a weak interaction and an elongation of X–N bond of B11XN12 nano-clusters. After the adsorption of CO gas over the doped positions of B11XN12 nano-cluster, the conductivity of the adsorbent and the atomic charges in some of the nearby B and N atoms around X atoms were dramatically enhanced. These calculations represent the capabilities of the B11XN12 nano-clusters in designing novel materials based on B11XN12 for potential applications in gas sensing.


1. Introduction

Carbon monoxide or CO is an extremely poisonous gas to human and animal bodies when encountered in concentrations above 35 ppm. It is a colourless, poisonous and odourless gas; these characteristics make it one of the most fatal gases with a significant role in air pollutants and in human daily life. Due to increasing concern over the safety and health hazards related to this gas, there is an increasing demand of CO sensors.1–3 Thus, theoretical studies regarding the adsorption and dissociation of CO molecule using the Rh-decorated SWCNT,4 Fe–B18N18 cluster,5 Al–graphitic boron nitride sheet,6 Al/Ga-doped BN nanotube,7 FeCo alloy (110) surface,8 ZnO (1010) surface,9 B12N12 nano-cage,10 doped h-BN monolayer,11 and transition metal doped BN nanotube12 have been reported. In particular, theoretical studies have shown that some metal-doped BN nanostructures are promising candidates to act as chemical sensors.13–17 During the last decade, various types of BN-based nanostructures consisting of nanotubes, nanohorns, nanoparticles, nanosheets, and nano-cages, have attracted considerable attention due to their unique chemical and physical properties.18–20 B12N12 is one of the known stable class of small III–V nano-clusters with the network of B–N bonds, which is energetically favorable. The 4-membered rings of B12N12 consist of the B–N bonds, which result in angular strain in the network. Since then, BiNi clusters have been widely investigated both theoretically and experimentally.21–26 Recently, Oku et al.27 synthesized B12N12, which can be detected by laser desorption time-of-flight mass spectrometry. Their study indicated that a BN nano-cluster is a structure built from four squares and six hexagons rings. In addition, Matxain and co-workers theoretically studied the electronic structure and the energy differences between covalent and van der Waals dimers using the hybrid B3LYP and MPW1PW91 density functional theory.28 Li et al. proposed bonding character and electronic structure of B12C6N6 fullerene using ab initio calculations.29 They found that replacing an N atom with a C atom in the B12N12 nano-cage results in a decrease of the cluster energy gap. Bahrami and co-authors reported a theoretical study of amphetamine adsorption upon the pure P-doped and Al-doped B12N12 nano-cage at the M06-2X/6-311++G** level of theory.30 Based on their results, the pure B12N12 nano-cage has a better condition for detecting amphetamine compared with the P-doped and Al-doped B11N12 nano-cage. Doping in BN nanomaterials can outstandingly alter the electronic properties, which can be utilized to develop its sensing applicability. Previous reports demonstrated the reaction of different molecules with the pure and doped metal BN nanostructures using DFT calculations.31–33 In this study, we investigated the interaction of CO molecule with the pure Ga, Ge, and Mg atoms of B11XN12 nano-clusters to find whether these adsorbents can be used as a gas sensor in environmental monitoring using DFT calculations.

2. Computational details

To study the structural parameters and electronic properties of the pure B11XN12 (X = Ge, Mg, and Ga) nano-cluster and that interacted with a CO molecule, we carried out density functional theory (DFT) calculations. The goal is to calculate equilibrium geometries, adsorption energies, Mulliken population charge analysis (MPA), molecular electrostatic potential (MEP), frontier molecular orbital (FMO), and density of states (DOS) analyses for these complexes. All the geometry optimizations and energy calculations were performed using GAMESS software34 at the level of density functional theory (DFT) using PBE functional augmented with an empirical dispersion term (PBE-D) and 6-311G** standard basis set. Moreover, the corresponding optical adsorption spectra were obtained through time-dependent density functional theory (TD-DFT) calculations in the gas phase regime.35–37 The relaxed structures were further subjected to the computations for harmonic vibrational frequencies at the PBE functional. For all the systems, the SCF convergence limit was set to 10−6 a.u. over energy and electron density. The basis set superposition error (BSSE) for the adsorption energy was corrected by implementing the counterpoise method.38,39 The adsorption energies (Ead) for each of the structures were determined at 298.15 K and 1 atm using the following formulae:
 
Ead = EB12N12–CO − (EB12N12 + ECO) + EBSSE (1)
 
Ead = EB11XN12–CO − (EB11XN12 + ECO) + EBSSE (2)
where EB11XN12–CO is the total energy of the B11XN12 nanocage interacting with the CO molecule. Enano-cage and ETM-nano-cage are the total energies of the pure and B11XN12, respectively, and ECO is the total energy of an isolated CO molecule. The electrophilicity index (ω) concept was stated for the first time by Parr et al.40,41 Chemical potential (μ) is defined according to the following equation:
 
μ = −χ =−1/2(I + A) (3)

Electronegativity (χ) is defined as the negative of chemical potential, as follows: χ = −μ. According to Koopmans theorem,42 chemical hardness (η) can be approximated by the following equation:

 
η = 1/2(IA) (4)

I(−EHOMO) is the ionization potential and (−ELUMO) is the electron affinity of the molecule. EHOMO is the energy of the highest occupied molecular orbital and ELUMO is the energy of the lowest unoccupied molecular orbital of the considered structure. Softness (S) and electrophilicity index are determined by the following equations, respectively:

 
S = 1/2η (5)
 
ω = (μ2/2η) (6)

3. Results and discussion

First, the B12N12 nano-cage was optimized as an adsorbent for detection of CO molecule at the gas phase using PBE-D method and 6-311G** standard basis set (see Fig. 1). Then, two stable states of the CO adsorption were considered, which locates the C and O over top of a B atom of B12N12 nano-cage. When CO molecule is adsorbed from C and O heads toward the boron atom of the adsorbent, the adsorption energies and distances of CO physisorption are −0.27 eV (II: 1.655 Å) and −0.06 eV (III: 3.176 Å) respectively (Table 1). Regarding previous studies, the adsorption of a CO molecule over (5, 5) SWCNT and capped CNT in their most stable states are about −0.147 and −0.02 eV.1,43 Baierlea et al. and Beheshtian et al. showed that the BN nanotubes have physical interactions with CO molecule in the range of about −0.13 and −0.06 eV.13,44 Our results were in agreement with the study of Beheshtian and co-workers, who reported a DFT study of CO molecule in reaction with B12N12 nano-cluster.10 They showed that CO can be adsorbed over the nano-cluster with the adsorption energy of 0.15–0.30 eV. It was found that the length of C[triple bond, length as m-dash]O bond (1.1386 Å at the PBE-D) after reaction with B12N12 is slightly smaller than that of a free CO molecule (1.357 Å), which is close to the reported experimental value of 1.128 Å.45 By Mulliken charge analysis, a charge transfer of about 0.172e (C head) and 0.019e (O head) occurs from CO molecule (electron donor) to B12N12 nano-cluster. In Fig. 1, the density of states are shown for the different CO-adsorptions over the B12N12 nano-cluster discussed above. Compared with the DOS of the pristine nano-cluster, the adsorption of CO molecule alters the electronic structure of B12N12, as shown in Fig. 1. This indicates a remarkable change (II) of energy gap (Eg) of about 45.8%. The DOS plots of the complex clearly reveal that the LUMO level has a distinct change after the adsorption process.
image file: c5ra12571e-f1.tif
Fig. 1 Optimized structures and density of state spectra of CO and B12N12 nano-clusters.
Table 1 Calculated bond length, adsorbate-surface distance D/Å, adsorption energy Ead/eV, HOMO energies (EHOMO/eV), LUMO energies (ELUMO/eV), dipole moment (DM/Debye), Fermi level energies (EF/eV) and HOMO–LUMO energy gap (Eg/eV) for the pure B12N12 nano-cluster
System Ead/eV D C–O/Å B–N/Å EHOMO/eV Eg/eV ELUMO/eV ΔEg% EF/eV DM/Debye
CO 1.1386 −8.87 7.03 −1.84 −5.36 0.226
I 1.493 −6.91 5.0 −1.91 −4.41 0.00
II −0.27 1.655 1.136 1.563 −6.52 2.71 −3.81 −45.8 −5.17 3.62
III −0.06 3.176 1.139 1.495 −6.90 4.81 −2.09 −3.80 −4.50 0.182


In this stage, first, the optimized structures and density of states of the perfect B12N12 nano-cluster with doped transition metal atoms containing Mg, Ge, and Ga are used to calculate the adsorption of a CO molecule, as shown in Fig. 2. The length of Mg–N, Ge–N, and Ga–N bonds of the doped nano-cluster are calculated with values of 2.079, 1.946, and 1.921 Å, respectively. In a similar study, Li et al. generated a similar structure by replacing N with C atom in the B12N12 nano-cage using ab initio calculations.29 Soltani et al.7 determined the length of a Ga–N bond of a B11XN12 nano-cage to be about 2.072 Å. As reported in Table 2, the angles of the N–Mg–N, N–Ge–N, and N–Ga–N bonds of B11XN12 nano-cluster were found to be about 69.66°, 76.92°, and 79.76°, respectively. As shown in Fig. 2, all the doped systems with Mg, Ge, and Ga atoms in B11XN12 nano-cluster belong to Cs point group, while the point group symmetry of B12N12 nano-cage is Th.46 In the B11XN12 nano-cluster, metal atoms move slightly out of the cluster form leading to a significant change of the local geometry of the nano-cluster. The point charges over the Mg, Ge, and Ga atoms in B11XN12 nano-cluster are 0.621, 0.720, and 0.739e, while the corresponding values for N atoms are about −0.522, −0.512, and −0.503e, respectively. To better understand the properties of crucial transition states for the X-doped B11XN12 nano-cluster, we plotted frontier molecular orbitals (HOMO and LUMO) for these systems. These orbitals play an important role in governing many chemical reactions, and they are also responsible for charge transfer properties (see Fig. 3). By comparison, it is found that the HOMO and LUMO energy levels can be significantly changed due to doping atoms because the energy gap of the nanoclusters reduces in the process of CO adsorption. From HOMO and LUMO analysis, it is easily seen that there is a uniform shift of the HOMO level to the upper energy region due to the mutual charge transfer and the interaction between N atoms of the nano-cluster and dopant atoms.


image file: c5ra12571e-f2.tif
Fig. 2 Optimized structures, infrared and density of state spectra of B11XN12 nano-clusters.
Table 2 The obtained structural parameters for a single CO adsorption over the pure and B11XN12 nano-clusters
System RC[double bond, length as m-dash]O RGa–N RGe–N RMg–N RN–Ga–N RN–Ge–N RN–Mg–N
GaB11N12 1.921      
State I 1.144 1.923 117.889
State II 1.133 1.938 114.828
State III 1.132 1.937 114.707
State IV 1.137 1.910 120.488
GeB11N12   1.946        
State I 1.164 1.922 107.243
State II 1.140 1.947 105.975
State III 1.137 1.934 107.591
State IV 1.138 1.890 105.814
MgB11N12     2.079      
State I 1.132 2.098 105.671
State II 1.138 2.064 105.079
State III 1.146 2.088 110.126
State IV 1.203 2.010 106.710



image file: c5ra12571e-f3.tif
Fig. 3 Crucial transition states of B11XN12 structures with the largest component coefficient marked.

The next stage involved search for a suitable nano-cluster for CO detection. After full optimization, adsorption states of CO molecule over the X-doped B12N12 nano-cluster using PBE method were studied and are reported in Table 3. This method has shown to obtain reasonable results for dative B–N bonds in comparison with other methods.47 In these interactions (Fig. 4–6), the CO molecule has electrostatic contact with the X-doped B12N12 nano-clusters. The observed trend for the B11GaN12 nano-cluster with CO molecule was found to be III > II > IV > I. In contrast, these values in the B11MgN12 were found to be III > I > II > IV, which means that the B11MgN12 is more stable than B11GaN12, whereas CO molecule is physisorbed toward B11GeN12 nano-cluster, which mainly stems from the van der Waals interaction, as can be seen in Table 3. According to the relative energy values, B11GeN12 is the least stable among all the species that interacted with the CO molecule. The minus value of adsorption energy in the interaction between adsorbent and adsorbate, representing that the CO molecule configuration has a very weak physical bond with the doping atoms of the nano-cluster. The weak type of interactions between adsorbate and adsorbent can be remarkable in gas detection because such ineffective interactions represent that the desorption of the adsorbate could be easy and the device can benefit from short recovery times.48 If Ead is significantly decreased, considerable shorter recovery time is expected. τ = υ0−1e(−Ead/kBT), where T is temperature, kB is the Boltzmann constant (8.62 × 10−5 eV K−1), and υ0 is the attempt frequency. According to this equation, the B11XN12 nano-cluster should be a good CO sensor with quick response as well as short recovery time. Upon the interaction of the CO molecule with X-doped B12N12 nano-cluster, the Mg–N, Ga–N, and Ge–N bond lengths shift to 2.098, 1.938, and 1.890 Å, respectively, with a resulting change in hybridization from sp2 to sp3. As a result, the adsorption of CO over the B11GaN12 and B11MgN12 nano-clusters leads to the elongation of the bond length but in the B11GeN12, it shortens the bond-length (Table 2). Bahrami and co-workers studied the effects of Al-doped and P-doped B12N12 nano-cage at the M06-2X method.30 They showed that the distance between the Al and the N atoms is about 1.97 Å. In addition, they reported that the adsorption of amphetamine on the surfaces of the perfect Al-doped and P-doped B12N12 nano-cages are found at −1.63, −2.48, and −0.33 eV, respectively. The results of these reports revealed that the Al-doped B11N12 can be used as an adsorbent in the environmental system.

Table 3 Calculated bond length, adsorbate-surface distance D/Å, adsorption energy Ead/eV, HOMO energies (EHOMO/eV), LUMO energies (ELUMO/eV), work function (Φ/eV), dipole moment (DM/Debye), Fermi level energies (EF/eV) and HOMO–LUMO energy gap (Eg/eV) for X-doped B11N12 nano-cluster
System Ead/eV D EHOMO/eV Eg/eV ELUMO/eV ΔEg% EF/eV Φ/eV DM/Debye
GaB11N12 −7.47 3.74 −3.73 −5.60 1.87 2.654
State I −0.200 2.511 −6.36 2.83 −3.53 24.33 −4.95 1.42 6.492
State II −0.657 2.148 −6.17 1.45 −4.72 61.23 −5.44 0.72 6.478
State III −0.668 2.148 −6.19 1.45 −4.70 60.16 −5.44 0.74 2.556
State IV −0.221 1.671 −6.28 5.023 −3.81 33.96 −5.04 1.23 4.562
GeB11N12 −6.38 4.41 −1.97 −4.18 2.21 1.673
State I −0.132 2.179 −4.423 2.88 −1.54 −34.74 −2.98 1.44 1.577
State II −0.067 3.476 −3.981 2.34 −1.64 −46.82 −2.81 1.17 1.188
State III −0.163 1.685 −6.288 4.93 −1.36 11.68 −3.83 2.47 1.8203
State IV −0.059 3.466 −6.270 4.73 −1.54 7.26 −3.91 2.37 1.840
MgB11N12 −6.10 3.14 −2.96 −4.53 1.57 5.249
State I −0.589 2.270 −5.86 1.44 −4.42 54.14 −5.14 0.72 7.246
State II −0.356 2.351 −5.90 4.741 −4.21 46.18 −5.0 0.79 5.38
State III −0.871 1.681 −5.92 2.93 −2.99 6.69 −4.45 1.46 6.85
State IV −0.073 1.358 −5.90 2.41 −3.49 23.25 −4.70 1.21 4.061



image file: c5ra12571e-f4.tif
Fig. 4 Optimized structures, infrared and density of state spectra of B11MgN12 nano-clusters interacting with a CO molecule.

image file: c5ra12571e-f5.tif
Fig. 5 Optimized structures, infrared and density of state spectra of B11GaN12 nano-clusters interacting with a CO molecule.

image file: c5ra12571e-f6.tif
Fig. 6 Optimized structures, infrared and density of state spectra of B11GeN12 nano-clusters interacting with a CO molecule.

To better understand the nature of CO adsorption over the electronic properties of adsorbent, we have carried out the electronic density of states (DOSs) analysis. The GaussSum program was used to obtain DOS results.49 The DOS spectrum of these structures represent the HOMO–LUMO gaps of the X-doped BN nano-clusters. Eg of Mg-doped, Ga-doped, and Ge-doped B12N12 nano-clusters are 3.14, 3.74, and 4.41 eV, while this value in the pure nano-cluster is about 5.0 eV. However, there is noticeable change in resistivity when the Eg for structures is decreased when compared to that of B12N12 nano-cluster. Oku and co-workers experimentally determined that the energy gap of the pure B12N12 nano-cluster is about 5.1 eV between HOMO and LUMO.27 Thus, the accuracy of the theoretical calculation is close to the experimental data as mentioned above. In the most stable configurations, after the CO adsorption toward the Mg, Ga, and Ge atoms of B11XN12 nano-clusters, ΔEg of these configurations significantly changed by about 54.14%, 61.23%, and 46.82%, respectively. This result implies that the electronic properties of the B11GaN12 nano-cluster are very sensitive to the CO adsorption in comparison with the pure B12N12 nano-cluster. Zhang et al. have shown that the graphitic BN sheet that by replacing Al-dopant and defects have high sensitive to the CO gas compared to the pure model.6 Moreover, their results indicated that the vacancy-defect g-BN has more change of energy gap than Al-dopant g-BN. As seen in Table 3, DOS spectrum in the CO/B11GaN12 complex (State II) near both of the valence and conduction levels have distinct changes in comparison with that in the B12N12. The valence level (−3.981 eV) of the CO/B11GeN12 complex (State II) exhibits a distinct change when compared to that of B12N12 (−6.38 eV). DOS spectrum for the CO/B11GeN12 complex (State I) revealed that the conduction level (−4.42 eV) of this complex exhibits a distinct change in comparison with that of B12N12 (−2.96 eV). This reports leads to a slight reduction in the work function that is important in the field emission applications. The values of Φ for the adsorption of CO gas, in the most stable models (Table 3), upon the Mg, Ga, and Ge atoms of B11XN12 nano-cluster are considerably decreased from 1.57, 1.87, and 2.21 eV to 0.72 (State I), 0.72 (State II), and 1.17 eV (State II), respectively. The decrement in the work function (Φ) values reveals that the field emission properties of the nano-clusters are more noticeable on the adsorption of CO gas because the electrons can be pulled from the surface more easily.50,51 In Fig. 7, the crucial excited states for the interaction of CO molecule with B11XN12 nano-cluster are provided, where the electron cloud in their LUMO, LUMO+3, and LUMO+5 are dominant over the CO molecule, while electron cloud in their HOMO orbitals do not have the same distribution of the electron density. To explore the interaction mechanism between CO and the B11XN12 nano-cluster, we studied molecular electrostatic potential (MEP) maps for these processes, where the positive charges over the X-doped B12N12 nano-cluster are represented by the blue colors (see Fig. 8). MEP plots are computed at an isovalue of 0.0004 e au−3. As demonstrated in these configurations, the CO molecule with blue color (positive charge) acts as an electron donor in the adsorption process. Most negative and natural regions of the MEP plots are colored yellow and green, respectively. Table 4 implies to the values of the quantum molecular descriptors computed for the adsorption of CO gas toward the B11XN12 nano-cluster. When CO gas reacts with the Mg-, Ga-, and Ge-doped B12N12 nano-cluster, the global hardness in the position of Ga and Mg dopants have the highest amount of hardness, while the lowest amount is in relation with the Ge dopant. In addition, when CO gas interacts with the B11XN12 nano-cluster, the η values have significant gain from 1.87, 1.57 and 2.21 eV in the Ga-, Mg-, and Ge-doped B12N12 nano-cluster to 2.54 eV (III), 2.59 eV (V), and 2.56 eV (V) after the adsorption of CO, respectively (see Table 4). For the suitable species, B11GaN12/CO, the global hardness of the complex is significantly increased and the electrophilicity of the complex is significantly reduced, indicating that the reactivity of the complex is decreased and also indicating that the stability of the complex is increased.52 The electrophilicity index of the Ga-, Mg- and Ge-doped B12N12 nano-clusters were higher than that of the CO molecule, suggesting a charge transfer from adsorbate to adsorbent.53–55


image file: c5ra12571e-f7.tif
Fig. 7 Crucial transition states of the B11XN12 structures interacting with CO with the largest component coefficient marked.

image file: c5ra12571e-f8.tif
Fig. 8 Calculated electrostatic potentials with a density isosurface of 0.0004 electrons au−3. The scale bar is in atomic units.
Table 4 Calculated quantum molecular descriptors for the CO molecule interacting with the B11XN12 nano-cluster
System I/eV A/eV η/eV μ/eV ω/eV χ/eV S/eV
GaB11N12 7.47 3.73 1.87 −5.6 8.38 5.6 0.27
State I 6.16 1.244 2.46 −3.70 2.79 3.70 0.20
State II 6.38 1.383 2.49 −3.88 3.02 3.88 0.20
State III 6.57 1.483 2.54 −4.03 3.18 4.03 0.196
State IV 6.310 1.287 2.51 −3.80 2.87 3.80 0.199
State V 6.482 1.197 2.64 −3.84 2.79 3.84 0.189
GeB11N12 6.38 1.97 2.21 −4.18 3.95 4.18 0.226
State I 4.423 1.545 1.44 −2.98 3.09 2.98 0.347
State II 3.981 1.636 1.17 −2.81 3.63 2.81 0.426
State III 6.288 1.363 2.46 −3.83 2.97 3.83 0.203
State IV 6.270 1.540 2.37 −3.91 3.22 3.91 0.211
State V 6.328 1.212 2.56 −3.77 2.78 3.77 0.195
MgB11N12 6.10 2.96 1.57 −4.53 6.53 4.53 0.318
State I 5.901 1.086 2.41 −3.49 2.53 3.49 0.207
State II 5.948 1.207 2.37 −3.58 2.70 3.58 0.210
State III 5.958 0.990 2.48 −3.47 2.43 3.47 0.201
State IV 6.178 1.318 2.43 −3.75 2.89 3.75 0.205
State V 6.077 0.903 2.59 −3.49 2.35 3.49 0.193


In addition, to investigate the stability of X-doped B12N12 nano-clusters, the calculation of the harmonic frequencies for all the adsorption models is needed. The vibrational frequencies of the Mg–N, Ga–N, and Ge–N in B11XN12 nano-clusters are at 590, 610 and 562 cm−1, respectively. IR vibrational spectrum indicates two active vibration modes (B–N bond) of B12N12 at 755 and 1375 cm−1 by the PBE-D method, which are in good agreement with the corresponding theoretical values of 1294 and 825 cm−1 by Pokropivny et al.56 and 1649 and 909 cm−1 by Jensen and Toftlund.57 Calculations over CO/B11MgN12 complex shows that the bond at 2190 cm−1 can be assigned to C–O molecule, while the values appearing at 1922 and 2159 cm−1 corresponds to the C–O stretching mode (from C head) of the CO adsorbed over B11GeN12 and B11GaN12 nano-clusters, respectively.

Herein, we use a TD-PBE/6-311G** calculation to study the optical properties of a CO molecule interacting with the B11XN12 nano-cluster.58 In Table 5, we have presented the lowest excitation modes of the four most stable configurations of the adsorbing CO systems. For pure B11GaN12 systems, we have two considerable peaks at the energies of 2.44 and 2.66 eV because they are related to the H−1− > L and H−2− > L vertical transitions. In this case, the H → L transition has very low intensity, which shows the wave functions of the HOMO and excitation modes to the adsorption state. As for I and IV states, the first excitation energy interval is between 2.54–2.85 eV, while for II and III states, the lowest excitation energy range is about 1.5 eV. In all the considered CO adsorbed states, the contribution of the H → L excitation has a noticeable growth. Similar trends also can be seen in corresponding Mg- and Ge-doped systems. Although the lowest excitation modes have a red shift in comparison with Ga-doped BN nanocage, in these cases, HOMO and LUMO have similar radial distribution on the cluster, which consequently results in low dipole moment and oscillation strength.

Table 5 Selected excitation energies (eV, nm), oscillator strength (f), and relative orbital contributions calculated with the PBE method
Methods Energy/eV Wavelength/nm Oscillator strength (f) Assignment
a H and L are the highest occupied (HOMO) and lowest unoccupied molecular orbitals (LUMO), respectively.
Ga doping 2.44 508 0.0040 H−1 → La (100%)
2.66 465 0.0070 H−2 → L (100%)
State I 2.85 435 0.0016 H → L+1 (100%)
2.90 428 0.0002 H−1 → L (100%)
State II 1.52 817 0.0084 H → L (99%)
1.54 804 0.0003 H−1 → L (99%)
State III 1.51 818 0.0035 H → L (96%)
1.53 806 0.0003 H−1 → L (99%)
State IV 2.54 488 0.0004 H−1 → L (100%)
2.82 439 0.0105 H−2 → L (99%)
Mg doping 0.66 1891 0.0010 H → L (100%)
1.17 1061 0.0006 H−1 → L (99%)
State I 0.71 1740 0.0011 H → L (100%)
1.22 1013 0.0008 H−1 → L (99%)
State II 0.70 1780 0.0012 H−1 → L (99%)
1.21 1024 0.0008 H−2 → L (64%)
State III 0.75 1646 0.0005 H−1 → L (99%)
1.12 1101 0.0007 H−2 → L (98%)
State IV 1.47 840 0.0016 H−1 → L (100%)
1.76 703 0.0020 H−2 → L (100%)
Ge doping 1.352 917 0.0020 H−1 → L (100%)
1.433 865 0.0014 H−2 → L (100%)
State I 1.973 628 0.0001 H → L (98%)
2.09 593 0.0032 H−1 → L (100%)
State II 1.46 848 0.0006 H−1 → L (100%)
1.61 772 0.0021 H−2 → L (100%)
State III 1.41 878 0.0020 H−1 → L (100%)
1.48 835 0.0011 H−2 → L (100%)
State IV 1.40 887 0.0020 H−1 → L (100%)
1.47 842 0.0012 H−2 → L (100%)


4. Summary

We have carried out density functional theory calculations over the adsorption of CO gas upon the Mg-, Ge-, and Ga-doped B12N12 nano-cluster. It was found that CO gas is weakly adsorbed over the B11XN12 nano-cluster due to van der Waals interaction between CO and the adsorbent. In addition, a slight reduction of the charge transfers in the adsorption process from CO to B11XN12 nano-cluster in comparison with the pure nano-cluster was observed. From the calculated results, we conclude that the electronic structure of the B11GaN12 undergoes a dramatic change after the CO adsorption process in comparison with the B11MgN12 and B11GeN12 nano-clusters. The B11GaN12 nano-cluster can act as a suitable sensor in the detection of CO gas by significantly changing the energy gap and work function of an adsorbent.

Acknowledgements

We thank the Sayad Shirazi Hospital, Golestan University of Medical Sciences, Gorgan, Iran. We also thank the Jaber Ebne Hayyan Unique Industry researchers Company, Gorgan city, Iran.

References

  1. K. Liao, V. Fiorin, D. S. D. Gunn, S. J. Jenkinsa and D. A. King, Phys. Chem. Chem. Phys., 2013, 15, 4059 RSC.
  2. S. Lin, X. Ye and J. Huang, Phys. Chem. Chem. Phys., 2015, 17(2), 888 RSC.
  3. C. Yeh and J. Ho, Phys. Chem. Chem. Phys., 2015, 17, 7555 RSC.
  4. J.-X. Zhao and Y.-H. Ding, Mater. Chem. Phys., 2008, 110, 411 CrossRef CAS.
  5. S. Nigam and C. Majumder, ACS Nano, 2008, 2, 1422 CrossRef CAS PubMed.
  6. Y.-H. Zhang, K.-G. Zhou, X.-C. Gou, K.-F. Xie, H.-L. Zhang and Y. Peng, Chem. Phys. Lett., 2010, 484, 266 CrossRef CAS.
  7. A. Ahmadi Peyghan, A. Soltani, A. Allah Pahlevani, Y. Kanani and S. Khajeh, Appl. Surf. Sci., 2013, 270, 25 CrossRef.
  8. P. Rochana and J. Wilcox, Surf. Sci., 2011, 605, 681 CrossRef CAS.
  9. Q. Yuan, Y.-P. Zhao, L. Li and T. Wang, J. Phys. Chem. C, 2009, 113, 6107 CrossRef CAS.
  10. J. Beheshtian, Z. Bagheri, M. Kamfiroozi and A. Ahmadi, Microelectron. J., 2011, 42, 1400 CrossRef CAS.
  11. S. Sinthika, E. Mathan Kumar and R. Thapa, J. Mater. Chem. A, 2014, 2, 12812 RSC.
  12. M. Bezi Javan, Surf. Sci., 2015, 635, 128 CrossRef.
  13. R. J. Baierlea, T. M. Schmidt and A. Fazzio, Solid State Commun., 2007, 142, 49 CrossRef.
  14. Z. Liu, Q. Xue, T. Zhang, Y. Tao, C. Ling and M. Shan, J. Phys. Chem. C, 2013, 117, 9332 CrossRef CAS.
  15. Q. Dong, X. M. Li, W. Q. Tian, X.-R. Huang and C.-C. Sun, J. Mol. Struct., 2010, 948, 83 CrossRef CAS.
  16. B. Gao, J.-X. Zhao, Q.-H. Cai, X.-G. Wang and X.-Z. Wang, J. Phys. Chem. A, 2011, 115, 9969 CrossRef CAS PubMed.
  17. S. Lin, X. Ye, R. S. Johnson and H. Guo, J. Phys. Chem. C, 2013, 117(33), 17319 CrossRef CAS.
  18. A. Nishiwaki, T. Oku and I. Narita, Sci. Tech. Adv. Mater., 2004, 5, 629 CrossRef CAS.
  19. T. Oku, K. Hiraga and T. Matsuda, Mater. Trans., 2008, 49, 2461 CrossRef CAS.
  20. C. Huang, W. Ye, Q. Liu and X. Qiu, ACS Appl. Mater. Interfaces, 2014, 6(16), 14469 Search PubMed.
  21. V. V. Pokropivny, V. V. Skorokhod, G. S. Oleinik, A. V. Kurdyumov, T. S. Bartnitskaya, A. V. Pokropivny, A. G. Sisonyuk and D. M. Sheichenko, J. Solid State Chem., 2000, 154, 214 CrossRef CAS.
  22. J. Guo-Qiang, C. Xin and H. Zheng, Chin. Phys. Lett., 2009, 26, 033601 CrossRef.
  23. A. V. Pokropivny, Diamond Relat. Mater., 2006, 15, 1492 CrossRef CAS.
  24. A. Soltani, M. T. Baei, M. Ramezani Taghartapeh, E. Tazikeh Lemeski and S. Shojaee, Struct. Chem., 2015, 26, 685 CrossRef CAS.
  25. A. Ahmadi Peyghan and H. Soleymanabadi, Curr. Sci., 2015, 108, 1910 Search PubMed.
  26. S. Yourdkhani, T. Korona and N. L. Hadipour, J. Phys. Chem. A, 2015, 119, 6446 CrossRef CAS PubMed.
  27. T. Oku, A. Nishiwaki and I. Narita, Sci. Tech. Adv. Mater., 2004, 5, 635 CrossRef CAS.
  28. J. M. Matxain, L. A. Eriksson, J. M. Mercero, X. Lopez, M. Piris, J. M. Ugalde, J. Poater, E. Matito and M. Sola, J. Phys. Chem. C, 2007, 111, 13354 CrossRef CAS.
  29. F. Li, Y. Zhang and H. Chen, Phys. E, 2014, 56, 216 CrossRef CAS.
  30. A. Bahrami, S. Seidi, T. Baheri and M. Aghamohammadi, Superlattices Microstruct., 2013, 64, 265 CrossRef CAS.
  31. N. Injan, J. Sirijaraensreb and J. Limtrakul, Phys. Chem. Chem. Phys., 2014, 16, 23182 RSC.
  32. N. L. Hadipour, A. Ahmadi Peyghan and H. Soleymanabadi, J. Phys. Chem. C, 2015, 119(11), 6398 CrossRef CAS.
  33. A. Soltani, M. T. Baei, M. Mirarab, M. Sheikhi and E. Tazikeh Lemeski, J. Phys. Chem. Solids, 2014, 75, 1099 CrossRef CAS.
  34. M. W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert, M. S. Gordon, J. H. Jensen, S. Koseki and N. Matsunaga, et al., J. Comput. Chem., 1993, 11, 1347 CrossRef.
  35. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  36. S.-l. Tang, Y.-j. Liu, H.-x. Wang, J.-x. Zhao, Q.-h. Cai and X.-z. Wang, Diamond Relat. Mater., 2014, 44, 54 CrossRef CAS.
  37. Q. Sun, Z. Li, D. J. Searles, Y. Chen, G. Lu and A. Du, J. Am. Chem. Soc., 2013, 135, 8246 CrossRef CAS PubMed.
  38. F. Tournus and J.-C. Charlier, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 71, 165421 CrossRef.
  39. M. T. Baei, M. Ramezani Taghartapeh, E. Tazikeh Lemeski and A. Soltani, Superlattices Microstruct., 2014, 72, 370 CrossRef CAS.
  40. R. G. Parr, R. A. Donelly, M. Levy and W. E. Palke, J. Chem. Phys., 1978, 68, 3801 CrossRef CAS.
  41. R. G. Parr, L. Szentpaly and S. Liu, J. Am. Chem. Soc., 1999, 121, 1922 CrossRef CAS.
  42. T. Koopmans, Physica, 1933, 1, 104 CrossRef CAS.
  43. Z. Li and C.-Y. Wang, Chem. Phys., 2006, 330, 417 CrossRef CAS.
  44. J. Beheshtian, M. T. Baei and A. Ahmadi Peyghan, Surf. Sci., 2012, 606, 981 CrossRef CAS.
  45. D. R. Lide, Handbook of Chemistry and Physics, CRC, Florida, 1992 Search PubMed.
  46. H. Y. Wu, X. F. Fan, J.-L. Kuo and W.-Q. Deng, Chem. Commun., 2010, 46, 883 RSC.
  47. M. T. Baei, Comput. Theor. Chem., 2013, 1024, 28 CrossRef CAS.
  48. J. Beheshtian, A. Ahmadi Peyghan and M. Noei, Sens. Actuators, B, 2013, 181, 829 CrossRef CAS.
  49. N. M. O'Boyle, A. L. Tenderholt and K. M. Langner, J. Comput. Chem., 2008, 29, 839 CrossRef PubMed.
  50. A. Ahmadi Peyghana, S. F. Rastegar and N. L. Hadipour, Phys. Lett. A, 2014, 378, 2184 CrossRef.
  51. J. Beheshtian, A. Ahmadi Peyghan and Z. Bagheri, Sens. Actuators, B, 2012, 171–172, 846 CrossRef CAS.
  52. Z. Bolboli Nojini and S. Samiee, J. Phys. Chem. C, 2011, 115, 12054 CrossRef.
  53. P. K. Chattaraj, U. Sarkar and D. R. Roy, Chem. Rev., 2006, 106, 2065 CrossRef CAS PubMed.
  54. R. G. Parr, L. V. Szentpály and S. Liu, J. Am. Chem. Soc., 1999, 121, 1922 CrossRef CAS.
  55. S. Armaković, S. J. Armaković, J. P. Šetrajčić, S. K. Jaćimovski and V. Holodkov, J. Mol. Model., 2014, 20, 2170 CrossRef PubMed.
  56. V. V. Pokropivny, V. V. Skorokhod, G. S. Oleinik, A. V. Kurdyumov, T. S. Bartnitskaya, A. V. Pokropivny, A. G. Sisonyuk and D. M. Sheichenko, J. Solid State Chem., 2000, 154, 214 CrossRef CAS.
  57. F. Jensen and H. Toftlund, Chem. Phys. Lett., 1993, 201, 89 CrossRef CAS.
  58. H. Ullah, A.-U.-H. Ali Shah, S. Bilal and K. Ayub, J. Phys. Chem. C, 2013, 117, 23701 CrossRef CAS.

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