Effects of Cl2 adsorption over the optical and electronic properties of Al12N12 and Al12CN11 fullerenes: Density functional theory study

Effects of Cl2 adsorption over the optical and electronic properties of Al12N12 and Al12CN11 fullerenes: Density functional theory study

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45 Contents lists available at ScienceDirect Physica E: Low-dimensional Systems ...

4MB Sizes 4 Downloads 39 Views

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

Contents lists available at ScienceDirect

Physica E: Low-dimensional Systems and Nanostructures journal homepage: www.elsevier.com/locate/physe

Effects of Cl2 adsorption over the optical and electronic properties of Al12N12 and Al12CN11 fullerenes: Density functional theory study

T

Fatemeh Azimi, Elham Tazikeh-Lemeski∗ Department of Chemistry, Gorgan Branch, Islamic Azad University, Gorgan, Iran

A R T I C LE I N FO

A B S T R A C T

Keywords: Cl2 Al12N12 Carbon doping Adsorption Optical properties DFT

Using the density functional theory (DFT) and time-dependent density functional theory (TD-DFT) calculations, we studied the adsorption and dissociation of Cl2 molecule over Al12N12 and Al12CN11 fullerenes at the room temperature. Chemisorption of Cl2 on the Al12CN11 (−4.41 eV) is much stronger in comparison with the Al12N12 fullerene (−2.92 eV). In the most stable states, we found that the gap energy of Al12N12 is significantly altered (ΔEg = 79%) upon the adsorption of Cl2 compared with Al12CN11 fullerene (ΔEg = 43%). UV–vis spectra analysis represents that the values of λmax are in the red region for the adsorption structures of Cl2 onto the pure and Al12CN11 fullerenes. We hope that our present theoretical studies can provide helpful information for further theoretical and experimental studies in the removal of this toxic gas by means of UV–vis and IR spectra.

1. Introduction Chlorine, Cl2, is a toxic greenish-yellow gaseous (occupational exposure limit (OEL) = 0.5 ppm) which is extensively applied for various industries related to plastics, paper products, dyestuffs, agrochemicals, pharmaceuticals, water purification and household cleaning products etc. Despite the useful various uses of chlorine, it has many harmful effects on human body like skin irritation, suffocation, sensory irritation, bronchospasm etc; and that is why there is an increasing demand of Cl2 sensors because of increasing concern over the safety and health hazards related to this gas [1–4]. Nowadays room temperature ppb level Cl2 sensing has been reported by using sulphonated copper phthalocyanine film with higher response and faster sensing characteristics as compared to the Sb-doped SnO2 nano-porous films, FEP/ polyaniline films and etc [5]. Altindal et al. have studied Cl2 sensing properties of crosswise-substituted phthalocyanines as a function of temperature (5–75 °C) in the concentration ranging from 50 to 150 ppb [6]. Beheshtian and his team have shown that the Cl2 molecule is strongly adsorbed on the ZnO nano-cluster via two mechanisms including chemisorption and dissociation with Gibbs free energy changes in the range of 0.36 to −0.92 eV at 298 K and 1 atm [7]. Nowadays, Al12N12 fullerene-like structures and nanotubes constructed of other elements have drawn considerable attention owing to their specific physical and chemical properties [8–12]. Particularly, fullerene-like nano-cages of group 3–5 elements in the periodic table have been theoretically identified and experimentally synthesized [13–19]. Especially the group III nitrides have been found as important source of ∗

Corresponding author. E-mail address: [email protected] (E. Tazikeh-Lemeski).

https://doi.org/10.1016/j.physe.2018.05.019 Received 10 February 2018; Received in revised form 15 May 2018; Accepted 18 May 2018

Available online 19 May 2018 1386-9477/ © 2018 Elsevier B.V. All rights reserved.

nanoscale materials because of their direct bond gaps affording optical and electro-optical properties and among these, AlN nanostructures had drawn considerable attention on the basis of ab initio calculation which shown in the large (AlN)n (n = 2–41) family, the Al12N12 nanocages are the most stable nanostructures energetically and thus it can be as an excellent candidate inorganic fullerene [20,21]. Many studies have been shown that the Al12N12 nano-cage may be used in the potential applications of hydrogen storage and gas sensor [12,22]. Jiao and coworkers recently studied the adsorption behavior of CO2 and N2 molecules over AlN single-walled nanotubes at the LDA and PW91 levels in a double numerical plus polarization basis set (DNP) [23]. They found that the tube diameter have a key effect on the chemisorption of CO2 respect to N2 molecule. Niu and co-workers introduced that the Al12N12 nano-cage by doping the alkali metal atom (Li, Na, and K) which can effectively reduce the energy gap values respect to the pure Al12N12 nano-cage [24]. Wang and co-workers reported potential of AlN nanostructures as hydrogen storage materials, while the binding of one H2 molecule is in the value of 0.1–0.2 eV which this binding leads to 4.7 wt % hydrogen storage [25]. Baei and co-authors have analyzed the OCN− chemisorption over AlN nanostructures by DFT simulation [26]. Soltani and co-workers have investigated the adsorption behavior of NO2 and SO2 molecules over Al12N12 nano-cage sensitized with gallium and magnesium using DFT calculations. They found that the Mg-doped Al11N12 nano-cage has high sensitivity to NO2 and SO2 molecules than Ga-doped Al11N12 nano-cage [27]. Recently, chigo Anota and coworkers have studied the adsorption, activation and possible dissociation of the glucose molecule on the magnetic [BN fullerene-B6]− system

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Fig. 1. Optimized structures of the Al12N12 and Al12CN11 fullerenes.

by means of density functional theory calculations [28]. In this work, we have taken the pure and Al12CN11 nano-cages as an efficient catalyst to remove of Cl2 and we evaluated the impact of adsorption on the structural, electronic, optical properties of the fullerenes.

Ead = EAl12N12 − Cl2 − (EAl12N12 + ECl2) + EBSSE

(1)

Ead = EAl12CN11 − Cl2 − (EAl12CN11 + ECl2) + EBSSE

(2)

where EAl12N12eCl2 and EAl12CN11eCl2 are the total energies of Al12N12 and Al12CN11 fullerenes interacting with Cl2. EAl12N12 and EAl12CN11 are total energies of the pure Al12N12 and Al12CN11 fullerenes. ECl2 is the total energy of Cl2 molecule. Natural bond orbital (NBO) analysis, molecular electrostatic potential (MEP), and total density of states (TDOS) analyses were calculated by using B3LYP method with 6–311+ +G** basis set.

2. Computational methods Geometry relaxations and total density of states (TDOS) analysis of the Cl2, Al12N12 and Al12CN11 fullerenes are performed with a DFTB3LYP method augmented with an empirical dispersion term (B3LYPD) [29,30] and then compared to B97 hybrid density functional accompanied by a dispersion-corrected GGA, B97-D [31]. All the geometrical relaxations and binding energies computations has been carried out using GAMESS suite of program [32] at the density functional theory (DFT) level using B3LYP-D/6–311++G** and B97-D/6–311+ +G** methods. The B3LYP-D and B97-D functionals of density functional theory have been considered by us and others to study the adsorption behavior and interaction details of gas molecules with the different nanostructures [33–36]. The basis set superposition error (BSSE) for the adsorption energy was corrected by using implementing the counterpoise method. The UV–vis absorption spectrum of Cl2 adsorbed over the Al12N12 and Al12CN11 surfaces have been carried out by means of time-dependent density functional theory (TD-DFT) calculations at the CAM-B3LYP functional and 6–311++G** basis set. We have determined the adsorption energy (Ead) of Cl2 on the pure and Al12CN11 fullerenes as follows:

3. Results and discussion 3.1. The Cl2 adsorbed on Al12N12 and Al12CN11 fullerenes Fig. 1 presents the relaxed structures of Al12N12 and Al12CN11 fullerenes that are made from 6 squares and 8 hexagons according to B3LYP-D and B97-D methods. NBO analysis demonstrates that the charge point value for Al and N atoms of Al12N12 are +0.109 and −0.109 e by B3LYP-D functional and +0.947 and −0.947 e by B97-D functional, respectively, while after doping process in the nano-cluster, the value of point charges for Al, C, and N atoms are found about +0.343, −0.433, and −0.169 e by B3LYP-D functional and +0.918, −0.857, and −0.947 e by B97-D functional, respectively. Based on natural bond orbital (NBO) analysis, the positive and negative charges upon Al and N atoms of Al12N12 fullerene is found to be +1.823 and −1.823 e, respectively, which are consistent with those reported by 36

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Fig. 2. Adsorption and dissociation of the Cl2 molecule on Al12N12 and Al12CN11 fullerenes.

theory exchange-correlation kernels for designing the electronic and optical properties of the nanostructures. All real frequencies of Cl2 interacting with Al12N12 fullerene reported here veritably indicate that these structures are true minima on their respective potential energy surfaces. Similar to recently study results, the adsorption and dissociation of chlorine (Cl2) onto the inorganic pure and Al12CN11 fullerenes were investigated in nine alternative states. As displayed in Tables 1 and 2, the bond lengths of Al12N12 and Al12CN11 fullerenes interacting with Cl2 molecule were calculated by means of B3LYP-D and B97-D functional. As shown in Tables 3 and 4, the calculated results demonstrated that all of the five states (I, II, III, IV, and V) exhibit notably large binding energy and distance interaction of −2.26 eV (2.149 Å), −2.99 eV (2.109 Å), −2.04 eV (2.165 Å), −0.76 eV (2.352 Å), and −0.79 eV (2.355 Å) by B3LYP-D functional, respectively. On the contrary, the adsorption and dissociation of chlorine

Saeedi and co-workers [37]. Al12N12 nano-cluster with Th symmetry has sufficiently high HOMO-LUMO energy gap (Eg) of 3.81 eV computed by B3LYP-D functional and 2.44 eV by B97-D functional. Based on calculated results, the HOMO-LUMO energy gap of Al12N12 fullerene is found about 3.93 eV (B3LYP) by Beheshtian and 3.92 eV by Baei [38,39]. The charge-density surfaces of HOMO (highest occupied molecular orbital) and LUMO (lowest unoccupied molecular orbital) for Al12N12 indicate that the HOMO is localized onto the N atoms and the LUMO is uniform localized through Al and N atoms of the substrate with an energy values of −6.45 and −2.61 eV by B3LYP-D and −5.45 and −3.01 eV by B97D, respectively. We have considered here the five adsorption configurations for Cl2 interacting with Al12N12 fullerene by B3LYP-D and B97-D functional, which are depicted in Fig. 2. The Becke 3-parameter hybrid functional or B3LYP is one of the most extensively applied density functional 37

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Table 1 Computed the bond length of the nano-cages, and the adsorbed Cl2 molecule over the surface of Al12N12 fullerene. Property B3LYP-D Al12N12 I II III IV V B97-D Al12N12 I II III IV V

Al1eN2

Al4eN2

Al1eN5

AleCl

N5Al1N2

N2Al3N6

Al4N2Al1

1.853 1.783 1.929 1.864 1.844 1.832

1.79 1.907 1.859 1.777 1.789 1.795

1.791 1.764 1.822 1.952 1.796 1.784

– 2.149 2.109 2.165 2.352 2.355

125 131.71 109.65 112.91 125.23 126.99

124.97 131.71 109.67 126.12 124.6 123.34

113.52 120.19 115.33 121.87 125.23 112.83

1.864 1.857 1.853 1.952 1.859 1.843

1.804 1.78 1.839 1.826 1.805 1.812

1.804 1.816 1.803 1.923 1.813 1.807

– 2.159 2.121 2.173 2.354 2.452

124.94 123.78 122.16 113.11 125.44 127.14

124.96 124.69 122.16 126.24 124.89 123.65

113.61 118.00 125.43 122.19 117.24 113.17

Table 2 Computed the bond length of the nano-cages, and the adsorbed Cl2 molecule over the surface of Al12CN11 fullerene. Property B3LYP-D Al12CN11 IX IIX IIIX IVX B97-D Al12CN11 IX IIX IIIX IVX

N1eAl2

Al2eC

Al6eC

AleCl

Al2CAl6

Al5N1Al2

N4Al2C

1.860 1.929 2.137 1.839 1.959

1.969 2.006 1.995 3.062 1.917

1.920 1.938 1.918 1.953 1.919

– 2.118 2.133 2.112 2.135

111.51 127.45 102.98 102.25 108.42

115.84 133.79 119.24 115.22 113.23

124.32 101.19 108.85 103.21 129.4

1.877 1.951 1.928 1.856 1.865

1.981 2.032 2.027 3.090 1.969

1.936 1.953 1.965 1.970 1.953

– 2.145 2.146 2.125 2.146

111.56 123.29 122.8 101.75 111.67

115.87 129.68 128.5 115.24 123.42

124.4 105.75 105.97 103.99 125.13

Table 3 Computed the binding energy and electronic properties of pure and carbon-doped Al12N12 fullerenes interacting with Cl2 molecule. Property B3LYP-D Al12N12 I II III IV V B97-D Al12N12 I II III IV V

Ead/eV

D/Ǻ

HOMO/eV

LUMO/eV

Eg/eV

ΔEg (%)

EF/eV

DM/Debye

– −2.15 −2.72 −1.87 −0.65 −0.73

– 2.149 2.109 2.165 2.352 2.355

−6.45 −6.07 −6.73 −6.50 −6.66 −6.65

−2.61 −3.29 −3.11 −5.71 −3.61 −3.61

3.84 2.78 3.62 0.79 3.05 3.50

– 27.60 5.73 79.42 20.57 8.85

−4.53 −4.68 −4.92 −6.11 −6.86 −6.86

0.0 9.73 7.04 0.65 5.50 5.47

– −2.45 −2.83 −2.19 −0.91 −1.12

– 2.15 2.12 2.165 2.35 2.34

−5.45 −5.14 −5.7 −5.10 −5.68 −5.69

−3.01 −3.99 −3.8 −4.60 −4.38 −4.4

2.44 1.15 1.90 0.50 1.3 1.29

– 52.87 22.13 79.51 46.72 47.13

−4.23 −4.56 −4.75 −4.85 −5.03 −5.04

0.0 9.21 7.01 0.72 6.43 6.51

Table 4 Computed the binding energy and electronic properties of Al12CN11 fullerenes interacting with Cl2 molecule. Property B3LYP-D Al12CN11 IX IIX IIIX IVX B97-D Al12CN11 IX IIX IIIX IVX

Ead/eV

D/Ǻ

HOMO/eV

LUMO/eV

Eg/eV

ΔEg (%)

EF/eV

DM/Debye

– −3.29 −2.70 −4.41 −0.79

– 2.12 2.13 2.11 2.14

−6.34 −6.55 −6.38 −6.23 −6.38

−2.54 −4.42 −2.81 −2.73 −2.87

3.80 2.13 3.57 3.86 3.51

– 43.94 6.05 1.57 7.63

−4.44 −5.485 −4.595 −4.48 −4.625

0.589 9.11 5.806 4.308 5.635

– −3.42 −2.98 −4.35 −0.95

– 2.09 2.11 2.07 2.12

−5.35 −6.05 −5.94 −5.28 −6.07

−2.98 −4.84 −5.19 −3.17 −4.85

2.37 1.21 0.75 2.11 1.22

– 48.94 68.35 10.97 48.52

−4.16 −5.44 −5.56 −4.22 −5.45

0.47 8.89 9.08 4.34 8.90

38

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Fig. 3. Frontier molecular orbitals in the most stable configurations.

significantly lengthened by the adsorption process from 1.853 and 1.790 Å to 1.929 and 1.859 Å, respectively. Compared to Al12N12, the bond angles of N5eAl1eN2 and N2eAl3eN6 after adsorption of chlorine can be significantly reduced from 125.0 and 124.97° to 109.65 and 109.67° which impel a local structural deformation in an adsorbent with an alteration from sp2 to sp3 hybridization. Our results demonstrated that the average bond length of CleCl in the Cl2 molecule (B3LYP-D: 2.049 Å and B97-D: 2.068 Å) is significantly elongation compared with that in the state IV (B3LYP-D: 2.777 Å and B97-D: 2.341 Å) and the state V (B3LYP-D: 2.240 Å and B97-D: 2.299 Å). Liu and Andryushechkin have shown that the CleCl bond length in the isolated Cl2 molecule is about 1.995 and 1.990 Å [40,41], which is in the good agreement with our calculation (2.049 Å by B3LYP-D and 2.068 Å by B97-D). Based on MPA analysis, the point charges on

upon inorganic Al12N12 fullerene is somewhat larger by B97-D than B3LYP-D functional (see Tables 3 and 4). As can be seen, the binding energy of chlorine on the fullerene in states I, II, III, IV, and V are −2.45, −2.99, −2.19, −1.09, and −1.12 eV by B97-D functional, respectively. Our results imply only when chlorine decomposed onto the nitrogen and aluminum atoms in the pure fullerene (state II) yield the memorable adsorbate-adsorbent interaction and indicated the interaction process is chemisorption in nature. Liu et al. indicated that the interaction between Cl2 molecule and the graphene surface is physisorption with the largest binding energy of 0.044 eV and the shortest interaction distance of 3.18 Å [40]. Beheshtian et al. reported that the values of adsorption and dissociation of chlorine over Zn12O12 fullerenes are in the range of −0.68 to −1.23 eV [7]. In the most stable configuration, the length of Al1eN2 and Al4eN2 bonds can be 39

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Fig. 4. Total density of states of the Cl2 molecule adsorbed on the Al12N12 and Al12CN11 fullerenes.

nitrogen and aluminum atoms are 0.835 and −1.209 e, respectively, while the point charges to the chlorine atom interacting with the aluminum atom of an adsorbent is about −0.417 e, indicating that the molecule function as an electron acceptor and the nano-cluster function as an electron donor. In the interim, the charge about 0.834 e transferring from the surface Al 3p valence orbital (Al12N12) to anti-bonding α* orbital of Cl2 molecule is the source of the Cl2 bond weakening (by B97-D in state II). Natural electron configuration calculations for Cl2 decomposed into Al atoms of the fullerene in state II reveals a valence electron configuration of 3s1.89 3p5.50 for the Cl2 and 3s0.39 3p0.82 for Al atom of the nano-cluster (natural bond orbital analysis was performed by using NBO 3.1 program at B3LYP/6–311++G**). Our calculations reveal that the hyper-conjugation between lone pair orbitals of aluminum as donors and some α* or π* orbitals of chlorine as acceptors can be happened. The most important common interaction in the Cl2/ Al12N12 (II) is LP(3)Cl → π* N4 to Al13. State II has the highest energy and

can make the complex more stable respect to other states. Also, the occupancy decreases to increasing the pz orbital contribution to the lone pair electrons of chlorine. For comparison, we investigate the effects of carbon doping upon the geometrical structure of the Al12N12 fullerene. Four stable formations of chlorine adsorbed onto Al12CN11 fullerenes are obtained, in which the fullerene is interacting with the Cl atoms of an adsorbate, as displayed in Fig. 2. After full optimization, the local symmetry of Al12CN11 fullerene is broken under a Jahn-Teller distortion [42,43] and indicated piddling deformation of the fullerene, because the ionic radii of carbon atom is close to that of the nitrogen atom of the nano-cluster [44]. The length of Al2eC and N1eAl2 bonds in the states IIIX increase from 1.969 and 1.579 Å to 3.062 and 1.839 Å, respectively. Nevertheless, when the Cl2 adsorbed onto Al12CN11 surface, the electronic configuration of an adsorbent represents sp3 hybridization and the bond lengths vicinal the adsorption site become longer and bond angles 40

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Fig. 5. The computed UV–vis absorption spectrum of the Al12N12 and Al12CN11, which is determined by a Gaussian functional convolution with FWHM = 3000 cm−1. The corresponding infrared spectrum of the Al12N12 and Al12CN11 fullerenes are shown as the insets.

charges of Al and C atoms are −0.430 and −0.896 eV, respectively. One of the most important factors in HOMO/LUMO interactions is the energy difference between HOMO of the Cl2 molecule (nucleophile agent) and that of LUMO of the Al12CN11 fullerene (electrophile agent). The computed energy levels and symmetry of HOMO, LUMO, as well as the HOMO-LUMO differences (Eg) between Cl2 and Al12CN11 fullerene is displayed in Table 4 and Fig. 3. The HOMO (α) and LUMO (α) energies of Al12N12 fullerene is calculated to be −6.45 and −2.61 eV by B3LYP-D functional and −5.45 and −3.01 eV by B97-D functional, respectively. Tahmasebi et al. showed the theoretical values of HOMO and LUMO energies of Al12N12 fullerene is found to be −6.49 and −2.65 eV by B3LYP functional [46]. Using B3LYP functional, Rad and Ayub have previously demonstrated the values of HOMO and LUMO of Al12CN11 fullerene is equal to −6.47 and −2.54 eV, respectively [47]. After the carbon doping on Al12N12 fullerene, the value of HOMO (α) and LUMO (α) energies are lowered by −6.34 and −2.54 eV, while the HOMO energy in the situation of beta (β) is lowered by −6.35 eV and LUMO energy in the situation of beta is increased to −3.48 eV, respectively. Our FMO analysis indicates that HOMO energy of the Cl2 molecule (nucleophile agent) is −8.69 eV, and LUMO energy for the Al12N12 and Al12CN11 fullerenes are −2.61 and −2.54 eV respectively (Table 4), suggesting that the large value of adsorption energy for Al12CN11 fullerene (Ead = −4.41 eV) may come from its lower LUMO energy level (−2.54 eV). Therefore, because of low energy level of LUMO of the Al12CN11 fullerene, it has a much stronger interaction with the Cl2 molecule. LUMO energy for the Al12N12 fullerene is −2.61 eV,

diminished, as the results are detailed in Tables 1 and 2. In Tables 3 and 4, the adsorption energies, the interaction distances, the charge transfers between Cl2 and Al12CN11 were studied. We found through calculations that there is a large increase in adsorption energy after Cl2 adsorbed on the surface of Al12CN11 compared to that of Cl2 adsorbed on Al12N12 fullerene, indicating the Cl2 endures a strong chemisorption on Al12CN11 fullerene. The most stable state, namely IIIX of Cl2 adsorbed onto Al12CN11 surface has adsorption energy of −4.52 eV with the interaction distance of 2.112 Å. Nevertheless, energetic favorability of the Cl2 adsorption onto the investigated fullerene is as follows: IIIX > IX > IVX > IIX. In this state, IIIX, we can see that the dissociation process is occurred for Cl2 molecule while one of Cl atom is situated on Al atom and other is situated on the C atom of an adsorbent as a result of the strong absorbability between the adsorbate and the Al12CN11, showing a strong chemisorption. However, the length of AleC bond in state IIIX increases when Cl2 adsorption happens to Cl atom and the length of AleN bond decreases when molecule interacts by means of Cl atom [45]. We can find that the dipole moment of Al12CN11 does change before and after Cl2 adsorption and these values have different behaviors for both functional. According to Table 4, the increment of DM can lead to the lowering of the adsorption energy and adsorption of Cl2 tremendously decreases the electronic properties of the systems. Small dipole moment (DM) for the state IIIX (4.31 Debye at B3LYP-D functional and 4.34 Debye at B97-D functional) causes significant increment in the adsorption energy and a little alteration in the conductivity of the system. The point charges for each Cl atom adsorbed on the Al and C atoms are found −0.111 and −0.639 eV, while the point 41

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Fig. 6. The computed UV–vis absorption spectrum of the Al12N12 and Al12CN11 fullerenes, which is determined by a Gaussian functional convolution with FWHM = 3000 cm−1.

suggesting that the lower value of adsorption energy (Ead = −2.92 eV) for the fullerene may come from its higher LUMO energy level. The results obtained from TDOS spectrum demonstrate that after doping of carbon instead of nitrogen atom in Al12N12 fullerene (see Fig. 4), Eg of the adsorbent finds a slight decline from 3.81 to 3.80 eV and EF of this system is slightly decreased from −4.53 to −4.44 eV [48]. In a study by Zope and Dunlap [49], the value of the energy band gap for Al24N24 fullerene was found to be 2.47 eV using the ΔSCF calculation. We found that the adsorption of Cl2 onto Al12N12 surface in the state III has the smallest Eg with the theoretical value of 0.79 and 0.50 eV by B3LYP-D and B97-D methods, respectively, which represents that the electronic property of the fullerene is more sensitive to Cl2 molecule by B97-D functional in comparison with B3LYP-D functional. Our calculations reveal that the distribution of Cl2 is situated between the valence and conduction bonds, which are nearby the EF. Based on TDOS plot, EF has obvious shift and the conductivity shift is plainly observable. In comparison with Al12N12 fullerene, Al12CN11 has less change in the region of Eg after adsorption of Cl2 molecule as the percentage change of Eg (ΔEg) in the most stable state (IX) is 35% less than

that of the state IV, as listed in Table 4. 3.2. Optical properties of Al12N12 and Al12CN11 fullerenes The UV–vis absorption spectrum of Cl2 adsorbed over the Al12N12 and Al12CN11 surfaces have been carried out by means of time-dependent density functional theory (TD-DFT) calculations at the CAM-B3LYP functional and 6–311++G** basis set as shown in Figs. 5 and 6. The excitation energies (ΔE), wavelength (λmax), and oscillator strengths (fosc) of main absorption spectrum are summarized in Table 5. To continue we explain excited states by delivering the complexes response to the oscillating electric field respect to a series of coupled occupied to virtual orbital excitations, which are mixed by suitable coefficients to deliver the demanded excited state. Table 5 represents the comparison between Al12N12 and Al12CN11 fullerenes before and after the adsorption of Cl2 molecule. We found the maximum excitation energy of 3.118 eV for Al12N12 fullerene by CAM-B3LYP functional. After adding of C atom to the Al12N12, the excitation energy for this the semiconductor nano-cage is reduced from 3.118 eV to 1.328, 1.744, and 42

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Table 5 Selected excitation energies (E, nm), oscillator strength (f), and relative orbital contributions of calculated at the CAM-B3LYP method. Methods

Energy/eV

λmax/nm

f

Assignment

Al12N12

4.676

265.2

0.0438

4.678

265.1

0.0435

3.187 3.314 3.631 3.818 3.299 3.796 3.974 0.832 1.032

389.9 374.1 341.4 324.7 375.8 326.5 311.9 489.9 500.9

0.0006 0.0027 0.0003 0.0009 0.009 0.001 0.001 0.0003 0.0008

1.057 3.079

472.3 402.6

0.0007 0.0011

3.160

392.3

0.0026

3.532 3.083 3.165

350.9 402.1 391.7

0.0053 0.0011 0.0026

3.534

350.8

0.0051

H-5 → L+1 (17%), H-3 → L+2 (15%), H-1 → L+4 (12%) H-5 → L+2 (12%), H-4 → L+3 (14%), H-2 → L+5(13%) H → L (98%) H → L (86%) H-2 → L (38%), H-1 → L (43%) H-2 → L (40%), H-1 → L (−34%) H-5 → L (77%) H-2 → L (−15%), H-1 → L (74%) H-3 → L+1 (−22%), H → L (69%) H → L (97%) H-11 → L (36%), H-5 → L (18%), H1 → L (18%) H-2 → L (98%) H-14 → L(30%), H-13 → L (12%), H-12 → L (29%) H-15 → L (16%), H-13 → L (47%), H-5 → L (15%) H → L (71%) H-14 → L (41%), H-12 → L (29%) H-15 → L (16%), H-13 → L (40%), H-5 → L (15%) H → L (72%)

3.697 2.914

335.4 425.4

0.0054 0.0181

2.460 0.937 1.262

503.9 985.1 952.3

0.0053 0.008 0.0053

1.362

909.8

0.0035

State IIX

1.403

783.7

0.0003

State IIIX

2.022 2.831 1.932 3.449 3.615

613.2 437.9 641.7 359.4 342.9

0.0001 0.0114 0.0001 0.0014 0.0015

State IVX

1.407

880.6

0.0003

2.019 2.841

613.9 436.3

0.0001 0.0109

State I

State II

State III

State IV

State V

Al12CN11

State IX

H-7 → L (85%), H-6 → L+1 (13%) H-7 → L (88%), H → L (2%), H10 → L (2%) H-5 → L (10%), H-4 → L (87%) H-5 → L (10%), H-5 → L+1(18%), H-4(A) → L (−13%), H-3 → L (25%) H-5 → L (10%), H-5 → L+1(18%), H → L (15%) H-2 → L (39%), H-1 → L (23%), H → L (17%) H-5 → L (68%) H-1 → L (19%), H → L (67%) H-2 → L (82%) H → L (79%) H → L (41%), H → L+1(25%), H1 → L (14%) H-2 → L (41%), H-1 → L (21%), H → L (18%) H-5 → L (69%) H-1 → L (20%), H → L (66%)

observed in the wavelengths of 438 (IIX) and 436 nm (IVX) with fosc of 0.0114 and 0.0109 as shown in Table 5.

2.119 eV in Al12CN11 fullerene. The strong absorption peak at the wavelength of 265.2 nm with fosc of 0.0438 for Al12N12 fullerene is observed using CAM-B3LYP functional. We also observed the wavelength of 405 nm with fosc of 0.0049 for Al12CN11 nano-cage at CAM-B3LYP functional. Tazikeh et al. demonstrated that the absorption spectrum of Al12N12 fullerene at the wavelengths of 399.17, 399.01, and 398.99 nm by B3LYP functional and 374.69, 374.57, and 374.53 nm by B3PW91 functional [48]. The data of UV–vis spectra indicate that the most significant changes in the optical properties of Al12N12 interacting with Cl2 related to the state IV. In the state IV, we have three noticeable absorption peaks at energies of 0.113, 0.346, and 0.566 eV because they are belong to the H → L (101%), H-2 → L (100%), and H-3 → L (99%) vertical transitions. Other remarkable contributions come from delocalized orbitals (H) or orbitals localized (L) over the Al12CN11/Cl2 complex (state IX) which indicate for each three optical transitions a low energy and are attributed to the combination of local excitations H → L (98%), H-3 → L (25%), and H-1 → L (63%). The maximum absorption wavelengths of Al12CN11/Cl2 complexes are in the visible spectra region, which are 985 and 880 nm in states IX and IVX, respectively. In other complexes, the maximum intensity absorption

3.3. Vibrational frequencies of Al12N12 and Al12CN11 fullerenes The theoretical infrared (IR) spectra of Cl2 adsorbed on Al12N12 and Al12CN11 fullerenes are illustrated in Figs. 5 and 7, in the regions 1600200 cm−1. All the vibration frequencies and thermodynamic parameters were calculated by means of B3LYP functional and 6–311+ +G** basis set. As can see in Fig. 7, Al12N12 fullerene demonstrates AleN stretching vibration at around 958 (I), 942 (II), 956 (III), 947 (IV), and 935 cm−1 (V) after Cl2 adsorption. The IR spectra for Al12N12 fullerene indicates AleN stretching vibration at 928 cm−1 which is close to the obtained results by Chang and co-workers [50,51]. For Al12CN11 fullerene, The AleC and AleN stretching vibrations observed in the regions of 863 and 925 cm−1, respectively. After adsorption of Cl2 on Al12CN11 fullerene, AleC and AleN stretching vibrations observed in the regions of 821 and 950 cm−1 (IX), 827 and 946 cm−1 (IIX), 728 and 970 cm−1 (IIIX), and 835 and 953 cm−1 (IVX). We observed the Cl2 decomposition on the carbon and aluminum atoms of 43

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

Fig. 7. The infrared spectrum of Cl2 adsorbed on the Al12N12 and Al12CN11 fullerenes.

depending on the orientation of Cl2 molecule upon the pure and Al12CN11 fullerenes.

Al12CN11 fullerene in the most stable state (IIIX) lead to the modes change and increase in bond strength with an adsorbent surface and simultaneous increment in bond length of AleC and AleN bonds agreeing recognized mode's changes to higher frequencies. The peaks of CeCl and AleCl in the most stable state (IIIX) appear at around 883 and 575 cm−1, while the peaks of NeCl and AleCl bonds appear at 671 and 534 cm−1, respectively.

References [1] A.K. Saroha, Safe handling of chlorine, J. Chem. Health Saf. 13 (2006) 5–11. [2] D. Wang, P. Hu, J. Xu, X. Dong, Q. Pan, Fast response chlorine gas sen-sor based on mesoporous SnO2, Sens. Actuators B 140 (2009) 383–389. [3] J. Tamaki, C. Naruo, Y. Yamamoto, M. Matsuoka, Sensing properties todilute chlorine gas of indium oxide based thin film sensors prepared by electron beam evaporation, Sens. Actuators B 83 (2002) 190–194. [4] F. Bender, C. Kim, T. Mlsna, J.F. Vetelino, Characterization of a WO3thin film chlorine sensor, Sens. Actuators B 77 (2001) 281–286. [5] R. Saini, A. Mahajan, R.K. Bedia, D.K. Aswal, A.K. Debnath, Room temperature ppb level Cl2 detection and sensing mechanism of highly selective and sensitive phthalocyanine nanowires, Sensor. Actuator B 203 (2014) 17–24. [6] A. Altindal, Z.Z. Öztürk, S. Dabak, Ö. Bekaroglu, Halogen sensing using thin filmsof crosswise-substituted phthalocyanines, Sensor. Actuator B 77 (2001) 389–394. [7] J. Beheshtian, A. Ahmadi Peyghan, Z. Bagheri, Adsorption and dissociation of Cl2 molecule on ZnO nanocluster, Appl. Surf. Sci. 258 (2012) 8171–8176. [8] M. Solimannejad, S. Kamalinahad, E. Shakerzadeh, Selective detection of toxic cyanogen gas in the presence of O2, and H2O molecules using a AlN nanocluster, Phys. Lett. A 380 (2016) 2854–2860. [9] A. Shokuhi Rad, K. Ayub, Adsorption properties of acetylene and ethylene molecules onto pristine and nickel-decorated Al12N12 nanoclusters, Mater. Chem. Phys. 194 (2017) 337–344. [10] A. Shokuhi Rad, K. Ayub, Coordination of nickel atoms with Al12X12 (X = N, P)

4. Conclusion In summary, we carried out a theoretical study upon the adsorption of Cl2 molecule onto the pure and Al12CN11 fullerenes by DFT calculations. Some important results are concluded as follows. (1) It was found that the adsorption of chlorine (Cl2) onto the pure and Al12CN11 fullerenes have strong covalent bond with each other as in this interaction each of the chlorine atoms sitting on an aluminum atom (see the states IV and IX). (2) The adsorption of Cl2 in the states IV and IX also damnably changed the electronic properties of the pure and Al12CN11 fullerenes by subtractive their energy gap and Fermi level. (3) Our study represents diverse optical and charge transfer properties 44

Physica E: Low-dimensional Systems and Nanostructures 103 (2018) 35–45

F. Azimi, E. Tazikeh-Lemeski

[11]

[12]

[13]

[14]

[15]

[16] [17] [18]

[19]

[20]

[21] [22] [23]

[24]

[25] [26]

[27]

[28]

[29] [30]

(1988) 785–789. [31] R. Peverati, K.K. Baldridge, J. Chem. Theor. Comput. 4 (12) (2008) 2030–2048. [32] M.W. Schmidt, K.K. Baldridge, J.A. Boatz, S.T. Elbert, M.S. Gordon, J.H. Jensen, S. Koseki, N. Matsunaga, K.A. Nguyen, S. Su, T.L. Windus, M. Dupuis, J.A. Montgomery, General atomic and molecular electronic structure system, J. Comput. Chem. 14 (1993) 1347–1363. [33] M. Bezi Javan, A. Soltani, Z. Azmoodeh, N. Abdolahi, N. Gholami, A DFT study on the interaction between 5-fluorouracil and B12N12 nanocluster, RSC Adv. 6 (2016) 104513–104521. [34] T. Umeyama, J. Baek, Y. Sato, K. Suenaga, F. Abou-Chahine, N.V. Tkachenko, H. Lemmetyinen, H. Imahori, Molecular interactions on single-walled carbon nanotubes revealed by high-resolution transmission microscopy, Nat. Commun. 6 (2015) 7732. [35] D. Tzeli, I.D. Petsalakis, G. Theodorakopoulos, Electronic structure and absorption spectra of supramolecular complexes of a fullerene crown ether with a p-extended TTF derivative, Phys. Chem. Chem. Phys. 13 (2011) 11965–11975. [36] M. Noei, Different electronic sensitivity of BN and AlN nanoclusters to SO2 gas: DFT studies, Vacuum 135 (2017) 44–49. [37] M. Saeedi, M. Anafcheh, R. Ghafouri, N.L. Hadipour, A computational investigation of the electronic properties of Octahedral AlnNn and AlnPn cages (n = 5 12, 16, 28, 36, and 48), Struct. Chem. 24 (2013) 681–689. [38] J. Beheshtian, A. Ahmadi Peyghan, Z. Bagheri, Selective function of Al12N12 nanocage towards NO and CO molecules, Comput. Mater. Sci. 62 (2012) 71–74. [39] M.T. Baei, M. Ramezani Taghartapeh, E. Tazikeh Lemeski, A. Soltani, Computational study of OCN- chemisorption over AlN nanostructures, Superlattice. Microst. 72 (2014) 370–382. [40] X.-Y. Liu, J.-M. Zhang, K.-W. Xu, Chlorine molecule adsorbed on graphene and doped graphene: a first-principle study, Physica B 436 (2014) 54–58. [41] B.V. Andryushechkin, V.V. Zheltov, V.V. Cherkez, G.M. Zhidomirov, A.N. Klimov, B. Kierren, Y. Fagot-Revurat, D. Malterre, K.N. Eltsov, Chlorine adsorption on Cu (111) revisited: LT-STM and DFT study, Surf. Sci. 639 (2015) 7–12. [42] M. Fujita, K. Wakabayashi, K. Nakada, K. Kusaka, Peculiar localized state at zigzag graphite edge, J. Phys. Soc. Jpn. 65 (1996) 1920–1923. [43] G.P. Soares, S. Guerini, Structural and electronic properties of impurities on boron nitride nanotube, J. Mod. Phys. 2 (2011) 857–863. [44] Y.-Y. Qi, Y. Zhang, J.-M. Zhang, V. Ji, K.-W. Xu, Structural and electronic properties of a single C chain doped zigzag AlN nanoribbon, Comput. Theor. Chem. 974 (2011) 151–158. [45] S.A. Aal, W.S. Abdel Halim, A.S. Shalabi, Cl2 adsorption on supported alkali metals and on the MgO and CaO (001) supports: a DFT study, Solid State Commun. 148 (2008) 464–468. [46] E. Tahmasebi, E. Shakerzadeh, Z. Biglari, Theoretical assessment of the electrooptical features of the group III nitrides (B12N12, Al12N12 and Ga12N12) and group IV carbides (C24, Si12C12 and Ge12C12) nanoclusters encapsulated with alkali metals (Li, Na and K), Appl. Surf. Sci. 363 (2016) 197–208. [47] A. Shokuhi Rad, K. Ayub, Detailed surface study of adsorbed nickel on Al12N12 nano-cage, Thin Solid Films 612 (2016) 179–185. [48] E. Tazikeh-Lemeski, Al12CN11 nano-cage sensitive to NH3 detection: a first-principles study, J. Mol. Struct. 1135 (2017) 166–173. [49] R.R. Zope, B.I. Dunlap, Electronic structure of fullerenelike cages and finite nanotubes of aluminum nitride, Phys. Rev. B 72 (2005) 045439. [50] C. Chen, A.B.S. Patzer, E. Sedlmayer, T. Steinke, D. Sulzle, Computational evidence for stable inorganic fullerene-like structures of ceramic and semiconductor materials, Chem. Phys. Lett. 350 (2001) 399–404. [51] C. Chen, A.B.S. Patzer, E. Sedlmayer, T. Steinke, D. Sulzle, A density functional study of small (AlN)X clusters: structure, energies, and frequencies, Chem. Phys. 271 (2001) 283–292.

nanocages enhances H2 adsorption: a surface study by DFT, Vacuum 133 (2016) 70–80. M.T. Baei, M. Ramezani Taghartapeh, E. Tazikeh Lemeski, A. Soltani, A computational study of adenine, uracil, and cytosine adsorption upon AlN and BN nanocages, Physica B 444 (2014) 6–13. A. Shokuhi Rad, K. Ayub, A comparative density functional theory study of guanine chemisorption on Al12N12, Al12P12, B12N12, and B12P12 nano-cages, J. Alloy. Comp. 672 (2016) 161–169. T. Oku, A. Nishiwaki, I. Narita, Formation and atomic structure of B12N12 nanocage clusters studied by mass spectrometry and cluster calculation, Sci. Technol. Adv. Mater. 5 (2004) 635–638. V.V. Pokropivny, V.V. Skorokhod, G.S. Oleinik, A.V. Kurdyumov, T.S. Bartnitskaya, A.V. Pokropivny, A.G. Sisonyuk, D.M. Sheichenko, Boron nitride analogs of fullerenes (the fulborenes), nanotubes, and fullerites (the fulborenites), J. Solid State Chem. 154 (2000) 214–222. J.M. Matxain, L.A. Eriksson, J.M. Mercero, X. Lopez, M. Piris, J.M. Ugalde, J. Poater, E. Matito, M. Sola, New solids based on B12N12 fullerenes, J. Phys. Chem. C 111 (2007) 13354–13360. A. Soltani, M.T. Baei, E. Tazikeh Lemeski, A.A. Pahlevani, Superlattic. Microstruct. 75 (2014) 716–724. M.D. Esrafili, R. Nurazar, Methylamine adsorption and decomposition on B12N12 nanocage: a density functional theory study, Surf. Sci. 626 (2014) 44–48. M.D. Esrafili, R. Nurazar, A density functional theory study on the adsorption and decomposition of methanol on B12N12 fullerene-like nanocage, Superlattice. Microst. 67 (2014) 54–60. W.-M. Sun, X.-H. Li, D. Wu, Y. Li, H.-M. He, Z.-R. Li, J.-H. Chen, C.-Y. Li, Theoretical study on superalkali-doped nanocages: unique inorganic electrides with high stability, deep-ultraviolet transparency, and considerable nonlinear optical response, Dalton Trans. 45 (2016) 7500–7509. Y. Zhang, X. Zheng, S. Zhang, S. Huang, P. Wang, H. Tian, Bare and Ni decorated Al12N12 cage for hydrogen storage: a first-principles study, Int. J. Hydrogen Energy 37 (2012) 12411–12419. J. Beheshtian, A. Ahmadi Peyghan, Z. Bagheri, Quantum chemical study of fluorinated AlN nano-cage, Appl. Surf. Sci. 259 (2012) 631–636. M.T. Baei, A. Soltani, S. Hashemian, H. Mohammadian, Al12N12 nanocage as a potential sensor for phosgene detection, Can. J. Chem. 92 (2014) 605–610. Y. Jiao, A. Du, Z. Zhu, V. Rudolph, S.C. Smith, A density functional theory study of CO2 and N2 adsorption on aluminium nitride single walled nanotubes, J. Mater. Chem. 20 (2010) 10426–10430. M. Niu, G. Yu, G. Yang, W. Chen, X. Zhao, X. Huang, Doping the alkali atom: an effective strategy to improve the electronic and nonlinear optical properties of the inorganic Al12N12 nanocage, Inorg. Chem. 53 (2014) 349–358. Q. Wang, Q. Sun, P. Jena, Y. Kawazoe, Potential of AlN nanostructures as hydrogen storage materials, ACS Nano 3 (3) (2009) 621–626. M.T. Baei, M. Ramezani Taghartapeh, E. Tazikeh Lemeski, A. Soltani, Computational study of OCN- chemisorption over AlN nanostructures, Superlattice. Microst. 72 (2014) 370–382. A. Soltani, S. Ghafouri Raz, M. Ramezani Taghartapeh, A. Varasteh Moradi, R. Zafar Mehrabian, Ab initio study of the NO2 and SO2 adsorption on Al12N12 nano-cage sensitized with gallium and magnesium, Comput. Mater. Sci. 79 (2013) 795–803. E. Chigo Anota, M. Salazar Villanueva, E. Shakerzadeh, M. Castro, Adsorption and possible dissociation of glucose by the [BN fullerene-B6]−magnetic nanocomposite. In silico studies, Appl. Nanosci. 26 (1) (2018) 52–60, http://dx.doi.org/10.1007/ s13204-018-0664-5. A.D. Becke, Density-functional thermochemistry, III. The role of exact exchange, J. Chem. Phys. 98 (1993) 5648–5654. T.C. Lee, W.T. Yang, R.G. Parr, Density-functional crystal orbital study on the structures and energetics of polyacetylene isomers, Phys. Rev. B Condens. Matter 37

45