This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Biaxial strain enhanced piezoelectric properties in monolayer g-C3N4\mathrm{C_{3}N_{4}}

San-Dong Guo1,2, Wen-Qi Mu1 and Yu-Tong Zhu1 1School of Electronic Engineering, Xi’an University of Posts and Telecommunications, Xi’an 710121, China 2Key Laboratary of Advanced Semiconductor Devices and Materials, Xi’an University of Posts and Telecommunications, Xi’an 710121, China
Abstract

Graphite-like carbon nitride (g-C3N4\mathrm{C_{3}N_{4}}) is considered as a promising candidate for energy materials. In this work, the biaxial strain (-4%-4%) effects on piezoelectric properties of g-C3N4\mathrm{C_{3}N_{4}} monolayer are studied by density functional theory (DFT). It is found that the increasing strain can reduce the elastic coefficient C11C_{11}-C12C_{12}, and increases piezoelectric stress coefficient e11e_{11}, which lead to the enhanced piezoelectric strain coefficient d11d_{11}. Compared to unstrained one, strain of 4% can raise the d11d_{11} by about 330%. From -4% to 4%, strain can induce the improved ionic contribution to e11e_{11} of g-C3N4\mathrm{C_{3}N_{4}}, and almost unchanged electronic contribution, which is different from MoS2\mathrm{MoS_{2}} monolayer (the enhanced electronic contribution and reduced ionic contribution). To prohibit current leakage, a piezoelectric material should be a semiconductor, and g-C3N4\mathrm{C_{3}N_{4}} monolayer is always a semiconductor in considered strain range. Calculated results show that the gap increases from compressive strain to tensile one. At 4% strain, the first and second valence bands cross, which has important effect on transition dipole moment (TDM). Our works provide a strategy to achieve enhanced piezoelectric effect of g-C3N4\mathrm{C_{3}N_{4}} monolayer, which gives a useful guidence for developing efficient energy conversion devices.

g-C3N4\mathrm{C_{3}N_{4}}, Piezoelectronics, 2D materials
pacs:
71.20.-b, 77.65.-j, 72.15.Jf, 78.67.-n                                    Email:[email protected]

I Introduction

Analogous to graphene, monolayer g-C3N4\mathrm{C_{3}N_{4}} has been achieved by top-down methods, which provides several potential applications including superior photocatalytic activitiesc2 ; c2-1 , sensingc3 and memory devicesc4 . For example, g-C3N4\mathrm{C_{3}N_{4}} can generate hydrogen from water under visible light with an appropriate band gap of 2.7 eVc5 . Many g-C3N4\mathrm{C_{3}N_{4}}/semiconductor heterostructures have been constructed to improve the photocatalytic performance of g-C3N4\mathrm{C_{3}N_{4}} to restrain the recombination of photogenerated carriers, like g-C3N4\mathrm{C_{3}N_{4}}/MoS2\mathrm{MoS_{2}}c6 , g-C3N4\mathrm{C_{3}N_{4}}/TiO2\mathrm{TiO_{2}}c7 and g-C3N4\mathrm{C_{3}N_{4}}/CdSc8 . Because of non-centrosymmetric structure, the monolayer g-C3N4\mathrm{C_{3}N_{4}} can exhibit a piezoelectricity, which may produce potential piezocatalysis applications.

In fact, due to potential nanoscale piezoelectric applications, the piezoelectricities of two-dimensional (2D) materials have attracted growing interestq4 . Experimentally, the piezoelectric coefficient (e11e_{11}=2.9×\times101010^{-10} C/m) of the monolayer MoS2\mathrm{MoS_{2}} has been measured with the 2H phaseq5 ; q6 , and an intrinsic vertical piezoelectric responseq8 has been proved to exist in the Janus MoSSe monolayer. The theoretical studies on piezoelectric properties of 2D materials, such as transition metal dichalchogenides (TMD), Janus TMD, group IIA and IIB metal oxides, group-V binary semiconductors and group III-V semiconductors, have been widely stuided q7 ; q7-1 ; q7-2 ; q7-3 ; q7-4 ; q9 ; q10 ; q11 ; q12 ; qr . The giant piezoelectricities in monolayer SnSe, SnS, GeSe and GeS have been reported, as high as 75-251 pm/Vq10 . A only in-plane piezoelectricity exits in many 2D materials, for example TMD monolayersq9 , and an additional out-of-plane piezoelectricity has also been predicted in many 2D Janus materialsq7 ; q7-2 ; q7-3 . A pure out-of-plane piezoelectric response has been predicted in penta-grapheneq7-4 , and two strategies are proposed to enhance its piezoelectric properties by strain and constructing Janus monolayer. The strain effects on the piezoelectric response of MoS2\mathrm{MoS_{2}}r1 , AsPq7-1 , SnSeq7-1 and Janus TMD monolayersr3 have been reported, and their piezoelectric properties can be effectively tuned. For example, the d22d_{22} of SnSe monolayer at -3.5% strain along the armchair direction is up to 628.8 pm/V from unstrained 175.3 pm/Vq7-1 .

Refer to caption
Figure 1: (Color online)The crystal structure of monolayer g-C3N4\mathrm{C_{3}N_{4}}, and the primitive cell is are marked by black line. The large red balls represent C atoms, and the small blue balls for N atoms. The rectangle supercell is marked by green line to calculate piezoelectric coefficients.

In this work, the biaxial strain-tuned piezoelectric properties of g-C3N4\mathrm{C_{3}N_{4}} monolayer are studied by using density functional perturbation theory (DFPT)pv6 . Only in-plane piezoelectricity exists for g-C3N4\mathrm{C_{3}N_{4}} monolayer. The independent piezoelectric constants d11d_{11} is predicted to be 1.42 pm/V. It is found that tensile strain of 4% can improve d11d_{11} to 6.12 pm/V, which is due to reduced C11C_{11}-C12C_{12} and enhanced e11e_{11}. Different from MoS2\mathrm{MoS_{2}}, increasing strain can enhance the ionic contribution to e11e_{11}. It is found that the gap of g-C3N4\mathrm{C_{3}N_{4}} increases from -4% to 4% strain. Strain can also induce the cross between the first and second valence bands, producing important effects on TDM. Therefore, our works give an experimental proposal to achieve enhanced piezoelectricity in g-C3N4\mathrm{C_{3}N_{4}} monolayer.

The rest of the paper is organized as follows. In the next section, we shall give our computational details and methods about piezoelectric coefficients. In the third section, we perform symmetry analysis for elastic and piezoelectric coefficients. In the fourth sections, we shall present main results of g-C3N4\mathrm{C_{3}N_{4}} monolayer. Finally, we shall give our conclusions in the fifth section.

Table 1: For monolayer g-C3N4\mathrm{C_{3}N_{4}}, the lattice constants a0a_{0} (Å\mathrm{{\AA}}), the elastic constants CijC_{ij} (Nm1\mathrm{Nm^{-1}}), shear modulus G2DG_{2D} (Nm1\mathrm{Nm^{-1}}), Young’s modulus C2DC_{2D} (Nm1\mathrm{Nm^{-1}}), Poisson’s ratio ν\nu, the HSE06 gaps (eV) and piezoelectric coefficients e11(sum)e_{11}(sum) [the electronic e11(ele)e_{11}(ele) and ionic e11(ion)e_{11}(ion) contribution] (101010^{-10} C/m ) and d11d_{11} (pm/V), with previous theoretical values and experimental results given in parentheses and square brackets.
a0a_{0} C11C_{11} C12C_{12} G2DG_{2D}
7.134 (7.135m1 )[7.130c5 ] 184.92 48.02 68.45
C2DC_{2D} ν\nu GapGap e11(sum)e_{11}(sum)
172.45 0.26 2.77(2.76m1 )[2.7c5 ] 1.94 (2.18nc )
e11(ele)e_{11}(ele) e11(ion)e_{11}(ion) d11d_{11}
3.02 -1.08 1.42
Refer to caption
Figure 2: (Color online) For monolayer g-C3N4\mathrm{C_{3}N_{4}}, the elastic constants (a) C11C_{11}-C12C_{12}, piezoelectric coefficients (b) e11e_{11} and (c) d11d_{11}, and (d) the ionic contribution and electronic contribution to e11e_{11} as a function of biaxial strain.

II Computational detail

Within the framework of DFT1 , we carry out our calculations by using the VASP codepv1 ; pv2 ; pv3 with the projected augmented wave (PAW) method. A kinetic cutoff energy of 500 eV is adopted, and we use the popular GGA of Perdew, Burke and Ernzerhof (GGA-PBE)pbe as the exchange-correlation potential to calculate piezoelectric and elastic properties. To avoid interactions between two neighboring images, a vacuum spacing of more than 19 Å\mathrm{{\AA}} along the z direction is added. The total energy convergence criterion is set to 10810^{-8} eV, and the Hellmann-Feynman forces on each atom are less than 0.0001 eV.Å1\mathrm{eV.{\AA}^{-1}}. The elastic stiffness tensor CijC_{ij} and the piezoelectric stress coefficients eije_{ij} are calculated by using strain-stress relationship (SSR) and DFPT methodpv6 . Within SSR and DFPT, the electronic and ionic contribution to the elastic and piezoelectric stress coefficients can be calculated directly from VASP code. The Brillouin zone sampling of g-C3N4\mathrm{C_{3}N_{4}} monolayer is done using a Monkhorst-Pack mesh of 11×\times11×\times1 for CijC_{ij}, and 6×\times11×\times1 for eije_{ij}. The Heyd-Scuseria-Ernzerhof (HSE06) hybrid functional with default parameters is used to obtain the correct electronic structures of monolayer g-C3N4\mathrm{C_{3}N_{4}}. The TDM are calculated by using VASPKIT codevk . The 2D elastic coefficients Cij2DC^{2D}_{ij} and piezoelectric stress coefficients eij2De^{2D}_{ij} have been renormalized by the the length of unit cell along z direction (LzLz): Cij2DC^{2D}_{ij}=LzLzCij3DC^{3D}_{ij} and eij2De^{2D}_{ij}=LzLzeij3De^{3D}_{ij}.

III Symmetry Analysis

In noncentrosymmetric crystals, a change of polarization can be induced by strain or stress. The phenomenon can be described by the third-rank piezoelectric stress tensors eijke_{ijk} and strain tensor dijkd_{ijk}, which are from the sum of ionic and electronic contributions:

eijk=Piεjk=eijkelc+eijkione_{ijk}=\frac{\partial P_{i}}{\partial\varepsilon_{jk}}=e_{ijk}^{elc}+e_{ijk}^{ion} (1)

and

dijk=Piσjk=dijkelc+dijkiond_{ijk}=\frac{\partial P_{i}}{\partial\sigma_{jk}}=d_{ijk}^{elc}+d_{ijk}^{ion} (2)

In which PiP_{i}, εjk\varepsilon_{jk} and σjk\sigma_{jk} are polarization vector, strain and stress, respectively. For 2D materials, if we only consider in-plane strain componentsq7 ; q9 ; q10 ; q11 ; q12 using Voigt notation, the dijd_{ij} can be derived by the relation:

(e11e12e16e21e22e26e31e32e36)=(d11d12d16d21d22d26d31d32d36)(C11C12C16C21C22C26C61C62C66)\left(\begin{array}[]{ccc}e_{11}&e_{12}&e_{16}\\ e_{21}&e_{22}&e_{26}\\ e_{31}&e_{32}&e_{36}\\ \end{array}\right)=\left(\begin{array}[]{ccc}d_{11}&d_{12}&d_{16}\\ d_{21}&d_{22}&d_{26}\\ d_{31}&d_{32}&d_{36}\\ \end{array}\right)\left(\begin{array}[]{ccc}C_{11}&C_{12}&C_{16}\\ C_{21}&C_{22}&C_{26}\\ C_{61}&C_{62}&C_{66}\\ \end{array}\right) (3)

The elastic tensor CijC_{ij} can be calculated by SSR, and the eije_{ij} can be attained by DFPT. The space group number of monolayer g-C3N4\mathrm{C_{3}N_{4}} is 187, and the corresponding point group 6¯m2\bar{6}m2 reduces eije_{ij}, dijd_{ij} and CijC_{ij} into:

(e11e11000e11000)\left(\begin{array}[]{ccc}e_{11}&-e_{11}&0\\ 0&0&-e_{11}\\ 0&0&0\\ \end{array}\right) (4)
(d11d110002d11000)\left(\begin{array}[]{ccc}d_{11}&-d_{11}&0\\ 0&0&-2d_{11}\\ 0&0&0\\ \end{array}\right) (5)
(C11C120C12C11000C11C122)\left(\begin{array}[]{ccc}C_{11}&C_{12}&0\\ C_{12}&C_{11}&0\\ 0&0&\frac{C_{11}-C_{12}}{2}\\ \end{array}\right) (6)

Here, the only in-plane d11d_{11} is derived by Equation 3:

d11=e11C11C12d_{11}=\frac{e_{11}}{C_{11}-C_{12}} (7)

IV Main calculated results

The structure of the monolayer g-C3N4\mathrm{C_{3}N_{4}} is illustrated in Figure 1, which shows a large aperture and a triazine ring as a unit. Firstly, the optimized lattice constants of monolayer g-C3N4\mathrm{C_{3}N_{4}} is aa=bb=7.134 Å\mathrm{\AA} using GGA, which agrees well with previous theoretical and experimental valuesm1 ; c5 . The band structure of the g-C3N4\mathrm{C_{3}N_{4}} monolayer using HSE06 is calculated, which shows an indirect band gap semiconductor with the valence band maximum (VBM) at Γ\Gamma point and the conduction band minimum (CBM) at K point. The position of CBM is different from previous one (at M point)m1 , which may be due to different HSE06 parameters. The calculated HSE06 band gap is 2.77 eV, which is very close to the experimental value (2.7 eV)c5 and the previous calculated result (2.76 eV)m1 . The independent elastic stiffness coefficients of C11C_{11} and C12C_{12} are calculated, and the monolayer has constants of C11C_{11}=184.92 Nm1\mathrm{Nm^{-1}} and C12C_{12}=48.02 Nm1\mathrm{Nm^{-1}}, which meet the Born criteria of mechanical stability. These elastic constants are larger than ones of MoS2\mathrm{MoS_{2}}q9 ; q11 . The 2D Young s moduli C2DC^{2D} and shear modulus G2DG^{2D} can be expressed asela :

C2D=C112C122C11C^{2D}=\frac{C_{11}^{2}-C_{12}^{2}}{C_{11}} (8)
G2D=C66G^{2D}=C_{66} (9)
C66=C11C122C_{66}=\frac{C_{11}-C_{12}}{2} (10)

The corresponding Poisson’s ratios is given:

ν2D=C12C11\nu^{2D}=\frac{C_{12}}{C_{11}} (11)

The calculated values are C2DC^{2D}=172.45 Nm1\mathrm{Nm^{-1}}, G2DG^{2D}=68.45 Nm1\mathrm{Nm^{-1}} and ν2D\nu^{2D}=0.26. The related data of g-C3N4\mathrm{C_{3}N_{4}} monolayer are listed in Table 1.

Refer to caption
Figure 3: (Color online)For monolayer MoS2\mathrm{MoS_{2}}, the elastic constants (a) C11C_{11}-C12C_{12}, piezoelectric coefficients (b) e11e_{11} and (c) d11d_{11}, and (d) the ionic contribution and electronic contribution to e11e_{11} as a function of biaxial strain.

The rectangle supercell is used to calculate piezoelectric stress coefficients of g-C3N4\mathrm{C_{3}N_{4}} monolayer by DFPT, and the x and y directions are shown in Figure 1. The calculated piezoelectric coefficient e11e_{11}=1.94×\times101010^{-10} C/m, being close to previous value e11e_{11}=2.18×\times101010^{-10} C/mnc . It is found that the electronic contribution is opposite to ionic contribution, and they are 3.02 ×\times101010^{-10} C/m and -1.08 ×\times101010^{-10} C/m, respectively. Based on calculated e11e_{11}, C11C_{11} and C12C_{12}, the predicted d11d_{11} is 1.42 pm/V, which is smaller than most 2D TMD monolayersq11 . Strain strategy is an effective method to improve piezoelectric effect of 2D materialsq7-1 ; r1 ; r3 . Here, we only consider biaxial strain, which can not produce polarization, not like uniaxial strain. In the simulation, the small biaxial strain (-4% to 4%) effects on piezoelectric properties of monolayer g-C3N4\mathrm{C_{3}N_{4}} are studied, which may be easily achieved in experiment. The elastic constants C11C_{11}-C12C_{12}, piezoelectric coefficients e11e_{11} and d11d_{11} as a function of biaxial strain are plotted in Figure 2. It is clearly seen that the C11C_{11}-C12C_{12} decreases, and e11e_{11} increases, when the strain changes from -4% to 4%. This will lead to improved d11d_{11} according to Equation 7 with 4% to 4% strain. At 4% strain, the d11d_{11} becomes 6.12 pm/V from unstrained 1.42 pm/V, increased by 331%.

The ionic contribution and electronic contribution to e11e_{11} as a function of biaxial strain are also plotted in Figure 2. It is found that there are narrow variations for electronic contribution with -4% to 4% strain, and only varies -0.35×\times101010^{-10} C/m. However, the magnitude change of the ionic contribution is very large, and about 4.52×\times101010^{-10} C/m. Therefore, the ionic contribution has an important role to enhance piezoelectric effect of g-C3N4\mathrm{C_{3}N_{4}} monolayer caused by strain, which is different from a typical 2D piezoelectric material MoS2\mathrm{MoS_{2}}. The MoS2\mathrm{MoS_{2}} monolayer has the same point group 6¯m2\bar{6}m2 with g-C3N4\mathrm{C_{3}N_{4}}, which gives rise to the same reduced piezoelectric coefficients. The elastic constants C11C_{11}-C12C_{12}, piezoelectric coefficients e11e_{11} and d11d_{11}, and the ionic contribution and electronic contribution to e11e_{11} of monolayer MoS2\mathrm{MoS_{2}} as a function of biaxial strain are plotted in Figure 3. For unstrained MoS2\mathrm{MoS_{2}}, our calculated C11C_{11} (131.76 Nm1\mathrm{Nm^{-1}}), C12C_{12} (31.20 Nm1\mathrm{Nm^{-1}}), e11e_{11} (3.78×\times101010^{-10} C/m) and d11d_{11} (3.76 pm/V) agree well with previous theoretical values (130 Nm1\mathrm{Nm^{-1}}, 32 Nm1\mathrm{Nm^{-1}}, 3.64×\times101010^{-10} C/m, 3.73 pm/V)q11 . For C11C_{11}-C12C_{12}, e11e_{11} and d11d_{11}, the change trend is similar to one of g-C3N4\mathrm{C_{3}N_{4}} with strain changing from -4% to 4%. However, the electronic part has positive contribution to improve piezoelectric effect of MoS2\mathrm{MoS_{2}}, while the ionic part gives negative effect. Thus, the electronic part dominate the enhancement of piezoelectric effect of monolayer MoS2\mathrm{MoS_{2}} caused by strain.

Refer to caption
Figure 4: (Color online)The energy band structures, TDM and energy band gap (Gap) of monolayer g-C3N4\mathrm{C_{3}N_{4}} using HSE06 with the application of biaxial strain (-4%-4%).

The monolayer g-C3N4\mathrm{C_{3}N_{4}} at applied strain, exhibiting piezoelectricity, not only should break inversion symmetry, but also should have a band gap. To confirm strained g-C3N4\mathrm{C_{3}N_{4}} to be a semiconductor, the energy band structures and gaps using HSE06 as a function of strain are plotted in Figure 4. It is clearly seen that the gap increases from 2.39 eV (-4%) to 3.07 eV (4%), but the positions of VBM and CBM do not change. It is found that the strain can induce the cross between the first and second valence bands at about 4% strain, which produces important effect on TDM. The TDM is the electric dipole moment associated with the transition between the two states, and we calculate the squares of TDM from the highest valence band to the lowest conduction band, which is also shown in Figure 4. The calculated results show that the outline of TDM has little change from -4% to 2%, which are mainly along Γ\Gamma-K and K-M. However, the magnitude of TDM becomes huge along Γ\Gamma-A and Γ\Gamma-B at 4% strain because of band cross between the first and second valence bands. So, strain can also produce important influence on optical absorptions of monolayer g-C3N4\mathrm{C_{3}N_{4}}.

V Conclusion

In summary, the reliable first-principles calculations are performed to investigate the biaxial strain (-4%-4%) effects on piezoelectric properties in monolayer g-C3N4\mathrm{C_{3}N_{4}}. Compared to unstrain one, compressive strain reduces e11e_{11}, and increases C11C_{11}-C12C_{12}. However, tensile strain produces opposite effects on e11e_{11} and C11C_{11}-C12C_{12}. These lead to improved d11d_{11} from compressive strain to tensile one. Calculated results show that the ionic contribution to e11e_{11} of g-C3N4\mathrm{C_{3}N_{4}} is in favour of the strain-induced enhanced d11d_{11}, which is different from MoS2\mathrm{MoS_{2}} monolayer. It is found that the HSE06 gap increases from 2.39 eV (-4%) to 3.07 ev (4%). The tensile strain (4%) can induce the cross between the first and second valence bands, which can induce huge TDM. Our predictive findings can provide a simple way to achieve energy-efficient energy transformation devices.

Acknowledgements.
This work is supported by the Natural Science Foundation of Shaanxi Provincial Department of Education (19JK0809). We are grateful to the Advanced Analysis and Computation Center of China University of Mining and Technology (CUMT) for the award of CPU hours and WIEN2k/VASP software to accomplish this work.

References

  • (1) P. Niu, L. L. Zhang, G. Liu and H. M. Cheng, Adv. Funct. Mater. 22, 4763 (2012).
  • (2) S. B. Yang, Y. J. Gong, J. S. Zhang, L. Zhan, L. L. Ma, Z. Y. Fang, R. Vajtai, X. C. Wang and P. M. Ajayan, Adv. Mater. 25, 2452 (2013).
  • (3) T. Y. Ma, Y. H. Tang, S. Dai and S. Z. Qiao, Small 10, 2382 (2014).
  • (4) F. Zhao, H. H. Cheng, Y. Hu, L. Song, Z. P. Zhang, L. Jiang and L. T. Qu, Sci. Rep. 4, 5882 (2014).
  • (5) X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and M. Antonietti, Nat. Mater. 8, 76 (2009).
  • (6) Y. Hou, A. B. Laursen, J. Zhang, G. Zhang, Y. Zhu, X. Wang et al., Angew. Chem., Int. Ed. 52, 3621 (2013).
  • (7) J. W. Zhou, M. Zhang and Y. F. Zhu, Phys. Chem. Chem. Phys. 17, 3647 (2015).
  • (8) J. Fu, B. B Chang,Y. L. Tian, F. N. Xi and X. P. Dong, J. Mater. Chem. A 1, 3083 (2013).
  • (9) W. Wu and Z. L. Wang, Nat. Rev. Mater. 1, 16031 (2016).
  • (10) W. Wu, L. Wang, Y. Li, F. Zhang, L. Lin, S. Niu, D. Chenet, X. Zhang, Y. Hao, T. F. Heinz, J. Hone and Z. L. Wang, Nature 514, 470 (2014).
  • (11) H. Zhu, Y. Wang, J. Xiao, M. Liu, S. Xiong, Z. J. Wong, Z. Ye, Y. Ye, X. Yin and X. Zhang, Nat. Nanotechnol. 10, 151 (2015).
  • (12) A. Y. Lu, H. Zhu, J. Xiao, C. P. Chuu, Y. Han, M. H. Chiu, C. C. Cheng, C. W. Yang, K. H. Wei, Y. Yang, Y. Wang, D. Sokaras, D. Nordlund, P. Yang, D. A. Muller, M. Y. Chou, X. Zhang and L. J. Li, Nat. Nanotechnol. 12, 744 (2017).
  • (13) L. Dong, J. Lou and V. B. Shenoy, ACS Nano, 11, 8242 (2017).
  • (14) S. D. Guo, X. S. Guo, Y. Y. Zhang and K. Luo, J. Alloy. Compd. 822, 153577 (2020).
  • (15) S. D. Guo, X. S. Guo, R. Y. Han and Y. Deng, Phys. Chem. Chem. Phys. 21, 24620 (2019).
  • (16) S. D. Guo, X. S. Guo, Z. Y. Liu and Y. N. Quan, J. Appl. Phys. 127, 064302 (2020).
  • (17) S. D. Guo and S. Q. Wang, J. Phys. Chem. Solids 140, 109375 (2020).
  • (18) M. N. Blonsky, H. L. Zhuang, A. K. Singh and R. G. Hennig, ACS Nano, 9, 9885 (2015).
  • (19) R. X. Fei, We. B. Li, J. Li and L. Yang, Appl. Phys. Lett. 107, 173104 (2015)
  • (20) K. N. Duerloo, M. T. Ong and E. J. Reed, J. Phys. Chem. Lett. 3, 2871 (2012).
  • (21) Y. Chen, J. Y. Liu, J. B. Yu, Y. G. Guo and Q. Sun, Phys. Chem. Chem. Phys. 21, 1207 (2019).
  • (22) Y. G. Guo, H. Q. Zhu and Q. Wang, ACS Appl. Mater. Interfaces 11, 1033 (2019).
  • (23) N. Jena, Dimple, S. D. Behere and A. D. Sarkar, J. Phys. Chem. C 121, 9181 (2017).
  • (24) Dimple, N. Jena, A. Rawat, R. Ahammed, M. K. Mohanta and A. D. Sarkar, J. Mater. Chem. A 6, 24885 (2018).
  • (25) X. Wu, D. Vanderbilt and D. R. Hamann, Phys. Rev. B 72, 035105 (2005).
  • (26) P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964); W. Kohn and L. J. Sham, Phys. Rev. 140, A1133 (1965).
  • (27) G. Kresse, J. Non-Cryst. Solids 193, 222 (1995).
  • (28) G. Kresse and J. Furthmu¨\ddot{u}ller, Comput. Mater. Sci. 6, 15 (1996).
  • (29) G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).
  • (30) J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • (31) V. Wang, N. Xu, J. C. Liu, G. Tang and W. T. Geng, arXiv:1908.08269v4 (2019).
  • (32) J. J. Liu, J. Phys. Chem. C 119, 28417 (2015).
  • (33) R. C. Andrew, R. E. Mapasha, A. M. Ukpong and N. Chetty, Phys. Rev. B 85, 125428 (2012).
  • (34) M. Zelisko, Y. Hanlumyuang, S. B. Yang, Y. M. Liu, C. H. Lei, J. Y. Li, P. M. Ajayan and P. Sharma, Nat. Commun. 5, 4284 (2014).