Skip to content
Publicly Available Published by De Gruyter November 6, 2015

A New Superhard Phase of C3N2 Polymorphs

  • Qun Wei EMAIL logo , Meiguang Zhang EMAIL logo , Haiyan Yan , Renxian Li , Xuanmin Zhu , Zhengzhe Lin and Ronghui Yao

Abstract

Carbon nitrides are excellent candidates for extreme hardness materials. In this work, a new I4̅3m phase of C3N2 has been uncovered by replacing part of the nitrogen atoms in the cagelike diamondoid nitrogen N10 with carbon atoms. This phase is mechanically and dynamically stable up to at least 50 GPa. The elastic anisotropy of I4̅3m-C3N2 is investigated by comparing with previously proposed α-C3N2. The tensile directional dependence of Young’s modulus obeys the following trend: E[111]>E[110]>E[100]. Electronic structure calculations reveal that I4̅3m-C3N2 is hole conducting. Hardness calculation shows that the I4̅3m-C3N2 is superhard with a hardness of 72.9 GPa.

1 Introduction

Searching new superhard material has been concentrated by researchers in recent years. Usually, borides, nitrides, and covalent of light elements (Be, B, C, N, etc.) are regarded as candidates of superhard materials [1–5]. Among these materials, carbon nitrides are a typical group. Theoretically, some superhard phases of carbon nitrides are predicted by first principles calculations before experimental preparation. Liu and Cohen [6, 7] presented a hexagonal superhard β-C3N4 structure, which has been obtained experimentally later [8, 9]. Then, C3N4 and C11N4 are theoretically proved to be superhard materials [10]. The theoretical hardness of C3N4 is 85.7 GPa [11], which is close to that of diamond. In addition, a group of tetragonal crystalline carbon mononitrides (CN) have been predicted [12]. By replacing part of the nitrogen atoms in the cubic gauche of nitrogen, a new superhard CN phase (cg-CN) has been proposed [13]. It is interesting that the cg-CN is found to be a metallic compound, although most superhard carbon nitrides are insulators or semiconductors. For example, Weihrich et al. proposed a Pa3̅ phase CN2 with a large bulk modulus of 405 GPa [14]. Using crystal structure prediction technique, a body-centered tetragonal CN2 has been uncovered [15]. The ideal strength and hardness calculation show that bct-CN2 is superhard with the hardness of 77.4 GPa. Tian et al. proposed Pm3̅m phase C3N2 (α-C3N2) at pressure below 1.0 GPa and P4̅3m phase (β-C3N2) at high pressure [16]. The Vickers hardness of both phases are approximately 86 GPa.

In the present work, we present a new metastable C3N2 phase by replacing part of the nitrogen atoms in the cagelike diamondoid nitrogen N10 [17] with carbon atoms. In N10 crystal, there are two inequivalent sites 12e and 8c. We replaced the N atoms in the Wyckoff position 12e (C3N2) and 8c (C2N3) with carbon atoms, respectively. Both C3N2 and C2N3 have the same space group I4̅3m. Phonon spectra calculations show that C2N3 is not stable, whereas C3N2 is found to be mechanically and dynamically stable up to at least 50 GPa, with a high hardness of 72.9 GPa.

2 Computational Details

Our calculations are performed using the VASP code [18] with the generalised gradient approximation (GGA) [19] and local density approximation (LDA) [20] for exchange correlation functional. The integration in the Brillouin zone is employed using the Monkhorst–Pack scheme (11×11×11), and the energy cut-off of 900 eV was chosen to ensure that energy calculations are converged to better than 1 meV/atom. The electron–core interactions are included by using the frozen-core all-electron projector augmented wave (PAW) method [21], with C:2s22p2 and N: 2s22p3 treated as the valence electrons. The elastic constants are determined from evaluation of stress tensor-generated small strain. Bulk modulus, shear modulus, Young’s modulus, and Poisson’s ratio are estimated by using Voigt–Reuss–Hill approximation [22].

3 Results and Discussions

After relaxing the crystal structure at 0 GPa, we obtain that this structure has a lattice parameter of 4.9202 Å within GGA. The optimised parameters for C3N2 at 0 GPa within GGA and LDA are listed in Table 1, and the unit cell structure is shown in Figure 1. In this structure, there are two inequivalent C6N4 cages. One is at body centered site, and the other is at the vertex site. Each cage connectes with six equivalent cages by C–C bonds, whereas there is no covalent bonding between center and vertex cages. The C6N4 cage is formed by 12 C–N bonds, and the bond angles of C–N–C and N–C–N are 110.0 and 106.4 degree, respectively. The P–V relations of α-C3N2, β-C3N2, and I4̅3m-C3N2 are plotted in Figure 2, as well as those of c-BN and diamond. We can see that the incompressibility of I4̅3m-C3N2 is larger than β-C3N2 and almost the same as α-C3N2 and c-BN. We can deduce that the bulk modulus of these three phases of C3N2 should follow the same order. To verify this, we obtain the values of equilibrium bulk modulus and its pressure derivative by fitting the E–V data at different pressure into the third-order Birch–Murnaghan equation of state (EOS) [23], and the results are listed in Table 1. The fitted bulk modulus for I4̅3m-C3N2 is 358 GPa within GGA, which is larger than that of β-C3N2 (328 GPa) and slightly less than that of α-C3N2 (373 GPa).

Table 1

Equilibrium lattice parameters (in Å), Wyckoff positions, volume of unit cell (in Å3), bonding length (in Å), EOS-fitted bulk modulus (in GPa) and its derivative for I4̅3m-C3N2.

CrystalMethoda0Wyckoff positionsV0dC–NdC–CB0B0
I4̅3m-C3N2GGA4.9202C(0.0, 0.0, 0.3537)119.11.49361.43943583.78
N(0.8281, 0.8281, 0.8281)
LDA4.8549C(0.0, 0.0, 0.3534)114.41.47291.42363983.71
N(0.8281, 0.8281, 0.8281)
α-C3N2GGA5.092132.01.43511.58863733.78
β-C3N2GGA5.088131.71.43141.58743282.98
1.4405
Figure 1: Crystal structure of I4̅3m-C3N2. The N atoms are represented in blue and C atoms black.
Figure 1:

Crystal structure of I4̅3m-C3N2. The N atoms are represented in blue and C atoms black.

Figure 2: Variation of ratio V/V0 with pressures for I4̅3m-C3N2, α-C3N2, β-C3N2, c-BN, and diamond.
Figure 2:

Variation of ratio V/V0 with pressures for I4̅3m-C3N2, α-C3N2, β-C3N2, c-BN, and diamond.

In order to check the stability of I4̅3m-C3N2, we calculate the elastic constants (see Tab. 2) and phonon frequencies (Fig. 3). For cubic crystal, the mechanical stability criteria at 0 GPa are as follows [24]: C11+2C12>0, C44>0, and C11C12>0. When a hydrostatic pressure is applied on crystals, the mechanical stability criteria for cubic crystals are as follows [25]: C11+2C12+P>0, C44P>0, and C11C12–2P>0. One can see from Table 2 that the calculated elastic constants are satisfied above mechanical stability criteria, which shows the mechanical stability of I4̅3m-C3N2 up to 50 GPa. The phonon dispersion curves for I4̅3m-C3N2 at 0 and 50 GPa are shown in Figure 3. No imaginary phonon frequency is observed in the whole Brillouin zone, indicating its dynamical stability in the range from 0 to at least 50 GPa. In the pressure range, we studied that the enthalpy of I4̅3m-C3N2 is always higher than that of α-C3N2 or β-C3N2, so it is a metastable phase of C3N2. The calculated bulk modulus (see Tab. 2) by means of the Voigh–Reuss–Hill approximation method is consistent with the EOS fitting results. The shear modulus quantifies its resistance to shear deformation and is a better indicator of potential hardness for ionic and covalent materials [26]. Young’s modulus E is defined as the ratio between stress and strain and is used to provide a measure of the stiffness of the solid. Compared to α-C3N2 and β-C3N2, the less E of I4̅3m-C3N2 means that α-C3N2 and β-C3N2 are stiffer than I4̅3m-C3N2. The G/B ratio can be used to determine the relative directionality of the bonding in the material. The calculated ratio G/B for I4̅3m-C3N2 is 0.64, which is smaller than that of α-C3N2 and β-C3N2 at the GGA level. This shows that the directionality of C–N bond in I4̅3m-C3N2 is weaker than that of α-C3N2 and β-C3N2.

Table 2

Calculated elastic constants Cij, bulk modulus B, shear modulus G, Young’s modulus E, and hardness Hk (in unit of GPa) of I4̅3m-C3N2.

CrystalMethodC11C12C44BGEνG/BHk
I4̅3m-C3N2GGA6532142403612325720.23550.6472.9
LDA6982512544002416040.24870.6076.8
α-C3N2GGA8561343203743357750.1550.9065.1
GGA [16]8761373273803658290.1360.96
β-C3N2GGA8501353213733357740.1540.9065.3
GGA [16]834983293433688130.1051.07

Also shown are Poission’s ratio ν and G/B ratio.

Figure 3: Phonon dispersion curve of I4̅3m-C3N2 at 0 GPa (a) and 50 GPa (b).
Figure 3:

Phonon dispersion curve of I4̅3m-C3N2 at 0 GPa (a) and 50 GPa (b).

Elastic anisotropy can give a prediction of the arrangement of the atoms in each direction, the bonding properties, and some chemical characters in different directions of materials. To illustrate the elastic anisotropy in detail, it is worthy to study the variation of Young’s modulus and shear modulus with direction. The variation of Young’s modulus along an arbitrary [hkl] direction for orthorhombic symmetry can be written as

(1)E1=s11(α4+β4+γ4)+(2s12+s44)(α2β2+β2γ2+α2γ2) (1)

where α, β, and γ are the direction cosines of the tensile stress direction, and s11, s12, and s44 are elastic compliance constants. The relations between sij and Cij are s11=(C11+C12)/(C112+C11C122C122),s12=C12/(C112+C11C122C122), and s44=1/C44. The directional dependence of Young’s modulus for I4̅3m-C3N2 is shown in Figure 4. β-C3N2 has almost the same elastic constants as α-C3N2 (see Tab. 2); thus, they should have the same anisotropy properties. So, we only illustrate the directional dependence of Young’s modulus for α-C3N2 in Figure 4 for comparison. It can be seen that the nonspherical nature of I4̅3m-C3N2 and α-C3N2 in Figure 4a and b shows the clear anisotropy of Young’s modulus. The maximum and minimum of Young’s modulus of I4̅3m-C3N2 in all directions are 589 and 547 GPa, respectively. The ratio Emax/Emin=1.08 is slightly less than that of α-C3N2 (Emax/Emin=820/747=1.10). The mean value of I4̅3m-C3N2 in all directions is 568 GPa, which matches well with the value calculated by the Voigt–Reuss–Hill approximation (572 GPa, see Tab. 2). In order to get a better understanding of the origin of the changes in Young’s modulus along different directions, the orientation dependences of Young’s modulus for I4̅3m-C3N2 and α-C3N2 are calculated when the tensile axis is within specific planes, as shown in Figure 5. From (1), we can obtain E−1=s11+sin22θ(2s12+s44–2s11)/4 for Young’s modulus in (100) plane, where θ is the angle between tensile stress and [001], and E−1=sin4θ(2s11+2s12+s44)/4+s11cos4θ+sin22θ(2s12+s44)/4 for Young’s modulus in (11̅0) plane, where θ is the angle between tensile stress and [001]. The maximum of Young’s modulus of I4̅3m-C3N2 is along [111] direction, whereas the minimum is along [100]. The order of Young’s modulus as a function of the principal crystal tensile for I4̅3m-C3N2 is E[111]>E[110]>E[100]. For α-C3N2, the ordering is E[100]>E[110]>E[111]. Using the Lyakhov–Oganov model [27], the hardness of I4̅3m-C3N2, as well as the hardness of α-C3N2 and β-C3N2, is calculated and listed in Table 2. The hardness of I4̅3m-C3N2 is 72.9 GPa within GGA. This shows that I4̅3m-C3N2 is a superhard material. Since C–C bond is a strong covalent bond, the shorter C–C bonds usually correspond to harder materials. For I4̅3m-C3N2, the C–C bond length is shorter than that of α-C3N2 and β-C3N2, so the hardness of I4̅3m-C3N2 is larger than that of α-C3N2 and β-C3N2.

Figure 4: Illustrations of directional dependence of Young’s modulus for I4̅3m-C3N2 (a) and α-C3N2 (b), and projections in xy plane of directional dependent Young’s modulus for I4̅3m-C3N2 (c) and α-C3N2 (d) at 0 and 50 GPa.
Figure 4:

Illustrations of directional dependence of Young’s modulus for I4̅3m-C3N2 (a) and α-C3N2 (b), and projections in xy plane of directional dependent Young’s modulus for I4̅3m-C3N2 (c) and α-C3N2 (d) at 0 and 50 GPa.

Figure 5: Plots of Young’s modulus for different crystallographic directions for I4̅3m-C3N2 (a) and α-C3N2 (b).
Figure 5:

Plots of Young’s modulus for different crystallographic directions for I4̅3m-C3N2 (a) and α-C3N2 (b).

To obtain a deeper insight into the hardness of the I4̅3m-C3N2, we calculate the band structure and density of state (DOS) and the atom resolved partial density of state (PDOS) of I4̅3m-C3N2 at 0 GPa, as shown in Figure 6. It is found that I4̅3m-C3N2 is metallic due to the finite electronic DOS at the Fermi level. The gap above fermi level shows that I4̅3m-C3N2 is a hole conductor, similar to BC5 [28] and BC7 [29]. From inspection of PDOS curves, it can be seen that C-p orbitals have a significant hybridisation with N-p orbitals near the Fermi level, signifying the strong C–N covalent bonding nature.

Figure 6: Band structure and DOS for I4̅3m-C3N2.
Figure 6:

Band structure and DOS for I4̅3m-C3N2.

4 Conclusions

A new I4̅3m-C3N2 phase has been uncovered by replacing nitrogen atoms of N10 with carbon atoms. Elastic and phonon calculations show that this phase is mechanically and dynamically stable up to at least 50 GPa. We have investigated the structural, mechanical, and electronic properties of I4̅3m-C3N2 using first principles calculations. The tensile directional dependence of Young’s modulus obeys the following trend: E[111]>E[110]>E[100]. The calculated results demonstrate that I4̅3m-C3N2 has large bulk and shear moduli. The hardness of 72.9 GPa illustrates its superhard character.


Corresponding authors: Qun Wei, School of Physics and Optoelectronic Engineering, Xidian University, Xi’an 710071, PR China, E-mail: ; and Meiguang Zhang, Department of Physics and Information Technology, Baoji University of Arts and Sciences, Baoji 721016, PR China, E-mail:

Acknowledgments

This work was supported by the Fundamental Research Funds for the Central Universities, the National Natural Science Foundation of China (Grant No. 11204007), the Natural Science Basic Research plan in Shaanxi Province of China (Grant Nos 2012JQ1005 and 2013JQ1007), and Education Committee Natural Science Foundation in Shaanxi Province of China (Grant No. 2013JK0638).

References

[1] Y. Tian, B. Xu, D. Yu, Y. Ma, Y. Wang, et al., Nature 493, 385 (2013).10.1038/nature11728Search in Google Scholar PubMed

[2] M. Zhang, Q. Wei, H. Yan, Y. Zhao, and H. Wang, J. Phys. Chem. C 118, 3202 (2014).10.1021/jp409152tSearch in Google Scholar

[3] X. Wang, J. Chem. Phys. 137, 184506 (2012).Search in Google Scholar

[4] Q. Wei, M. Zhang, L. Guo, H. Yan, X. Zhu, et al., Chem. Phys. 415, 36 (2013).Search in Google Scholar

[5] H. Yan, M. Zhang, Q. Wei, and P. Guo, Comput. Mater. Sci. 68, 174 (2013).Search in Google Scholar

[6] A. Liu and M. Cohen, Science 245, 841 (1989).10.1126/science.245.4920.841Search in Google Scholar PubMed

[7] A. Liu and M. Cohen, Phys. Rev. B 41, 10727 (1990).10.1103/PhysRevB.41.10727Search in Google Scholar PubMed

[8] K. Yu, M. Cohen, E. Haller, W. Hansen, A. Liu, et al., Phys. Rev. B 49, 5034 (1994).10.1103/PhysRevB.49.5034Search in Google Scholar

[9] C. Niu, Y. Z. Lu, and C. M. Lieber, Science 261, 334 (1993).10.1126/science.261.5119.334Search in Google Scholar PubMed

[10] M. Mattesini and S. Matar, Phys. Rev. B 65, 075110 (2002).10.1103/PhysRevB.65.075110Search in Google Scholar

[11] Y. Zhang, H. Sun, and C. Chen, Phys. Rev. B 76, 144101 (2007).10.1103/PhysRevB.76.085318Search in Google Scholar

[12] E. Kim, C. Chen, T. Köhler, M. Elstner, and T. Frauenheim, Phys. Rev. Lett. 86, 652 (2001).Search in Google Scholar

[13] X. Wang, K. Bao, F. Tian, X. Meng, C. Chen, et al., J. Chem. Phys. 133, 044512 (2010).Search in Google Scholar

[14] R. Weihrich, V. Eyert, and S. F. Matar, Chem. Phys. Lett. 373, 636 (2003).Search in Google Scholar

[15] Q. Li, H. Liu, D. Zhou, W. Zheng, Z. Wu, et al., Phys. Chem. Chem. Phys. 14, 13081 (2012).Search in Google Scholar

[16] F. Tian, J. Wang, Z. He, Y. Ma, L. Wang, et al., Phys. Rev. B 78, 235431 (2008).10.1103/PhysRevB.78.235431Search in Google Scholar

[17] X. Wang, Y. Wang, M. Miao, X. Zhong, J. Lv, et al., Phys. Rev. Lett. 109, 175502 (2012).Search in Google Scholar

[18] G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996).10.1103/PhysRevB.54.11169Search in Google Scholar PubMed

[19] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).Search in Google Scholar

[20] D. M. Ceperley and B. J. Alder, Phys. Rev. Lett. 45, 566 (1980).Search in Google Scholar

[21] G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).10.1103/PhysRevB.59.1758Search in Google Scholar

[22] R. Hill, Proc. Phys. Soc. London 65, 349 (1952).10.1088/0370-1298/65/5/307Search in Google Scholar

[23] F. Birch, Phys. Rev. 71, 809 (1947).Search in Google Scholar

[24] Z. Wu, E. Zhao, H. Xiang, X. Hao, X. Liu, et al., Phys. Rev. B 76, 054115 (2007).10.1103/PhysRevB.76.059904Search in Google Scholar

[25] J. Wang, J. Li, S. Yip, S. Phillpot, and D. Wolf, Phys. Rev. B 52, 12627 (1995).10.1103/PhysRevB.52.12627Search in Google Scholar PubMed

[26] J. M. Léger, P. Djemia, F. Ganot, J. Haines, A. S. Pereira, et al., Appl. Phys. Lett. 79, 2169 (2001).Search in Google Scholar

[27] A. Lyakhov and A. Oganov, Phys. Rev. B 84, 092103 (2011).10.1103/PhysRevC.84.055206Search in Google Scholar

[28] Q. Li, H. Wang, Y. Tian, Y. Xia, T. Cui, et al., J. Appl. Phys. 108, 023507 (2010).Search in Google Scholar

[29] H. Liu, Q. Li, L. Zhu, and Y. Ma, Solid State Commun. 151, 716 (2011).Search in Google Scholar

Received: 2015-6-16
Accepted: 2015-10-6
Published Online: 2015-11-6
Published in Print: 2015-12-1

©2015 by De Gruyter

Downloaded on 20.4.2024 from https://www.degruyter.com/document/doi/10.1515/zna-2015-0270/html
Scroll to top button