Next Article in Journal
Using Central Composite Experimental Design to Optimize the Degradation of Tylosin from Aqueous Solution by Photo-Fenton Reaction
Next Article in Special Issue
The Lateral Compressive Buckling Performance of Aluminum Honeycomb Panels for Long-Span Hollow Core Roofs
Previous Article in Journal
Engineering and Scaling the Spontaneous Magnetization Reversal of Faraday Induced Magnetic Relaxation in Nano-Sized Amorphous Ni Coated on Crystalline Au
Previous Article in Special Issue
Discrete Element Method Modeling of the Rheological Properties of Coke/Pitch Mixtures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two Novel C3N4 Phases: Structural, Mechanical and Electronic Properties

1
Key Laboratory of Ministry of Education for Wide Band-Gap Semiconductor Materials and Devices, School of Microelectronics, Xidian University, Xi’an 710071, China
2
School of Physics and Optoelectronic Engineering, Xidian University, Xi’an 710071, China
*
Author to whom correspondence should be addressed.
Materials 2016, 9(6), 427; https://doi.org/10.3390/ma9060427
Submission received: 26 April 2016 / Revised: 20 May 2016 / Accepted: 24 May 2016 / Published: 30 May 2016
(This article belongs to the Special Issue Computational Multiscale Modeling and Simulation in Materials Science)

Abstract

:
We systematically studied the physical properties of a novel superhard (t-C3N4) and a novel hard (m-C3N4) C3N4 allotrope. Detailed theoretical studies of the structural properties, elastic properties, density of states, and mechanical properties of these two C3N4 phases were carried out using first-principles calculations. The calculated elastic constants and the hardness revealed that t-C3N4 is ultra-incompressible and superhard, with a high bulk modulus of 375 GPa and a high hardness of 80 GPa. m-C3N4 and t-C3N4 both exhibit large anisotropy with respect to Poisson’s ratio, shear modulus, and Young’s modulus. Moreover, m-C3N4 is a quasi-direct-bandgap semiconductor, with a band gap of 4.522 eV, and t-C3N4 is also a quasi-direct-band-gap semiconductor, with a band gap of 4.210 eV, with the HSE06 functional.

1. Introduction

Studies on light element-based materials trace back to the middle of the last century. Since Lavoisier found that diamond was isostructural to carbon and much denser than graphite, many studies have been devoted to its synthesis under high pressure [1,2,3,4]. More and more researchers have begun to investigate the carbon allotropes [5,6,7,8,9,10,11,12,13,14,15]. The second light element-based material to be evidenced was boron nitride. It includes three different structures: blende-, wurtzite- and graphitic-type structures. Cubic boron nitride (c-BN) was first elaborated upon in 1957 by Wentorf, who performed direct conversion using graphitic boron nitride (at 7 GPa and 1500 °C) [16]. Many boron nitride allotropes have been investigated by researchers, such as O-BN, Pbca-BN, Z-BN, W-BN, h-BN, bct-BN, P-BN, and cT8-BN. Interest in carbon nitrides has been initiated by studying materials that exhibit mechanical properties comparable with those of diamond. A fullerene is a molecule of carbon in the form of a hollow sphere, ellipsoid, tube, and many other shapes. Gueorguiev et al. [17,18] studied the formation mechanisms and structural features of fullerene-like carbon nitride (FL CNx), utilizing first-principles calculations.
Liu et al. first predicted β-C3N4 [19]; its structure originated in β-Si3N4, with carbon substituting for silicon. In the same way, α-C3N4 has been deduced from α-Si3N4, replacing silicon with carbon. The bulk of α-C3N4 and β-C3N4 is 387 and 427 GPa, respectively, which are slightly smaller than that of diamond (431 GPa [20]). Therefore, there are sp2 and sp3 hybridizations on carbon and nitrogen in α-C3N4 and β-C3N4, respectively. The pseudocubic-C3N4 structure is isostructural to α-CdIn2Se4 [21] and was first proposed by Liu and Wentzcovitch [22]. The network structure of pseudocubic-C3N4 consists of corners-linked CN4 tetrahedra in which the C-N-C angle is close to 109°, which ensures sp3 hybridization for nitrogen. The bulk modulus of pseudocubic-C3N4 is 448 GPa, which is slightly larger than that of diamond.
Cubic-C3N4 is another C3N4 phase and is isostructural to the high-pressure structure of Zn2SiO4, which was proposed by Teter and Hemley [23]. The structure of cubic-C3N4 is similar to that of pseudocubic-C3N4, including the hybridization. Mo et al. [24] and later Kroll [25] proposed a γ-C3N4 polymorph derived from a γ-Si3N4 spinel high-pressure structure. The largest difference between these structures involves the hybridization of nitrogen and carbon. In pseudocubic-C3N4 or cubic-C3N4, both carbon and nitrogen adopt sp3 hybridization. Graphite C3N4 (g-C3N4) consists of the stacking along the c-axis of graphitic planes. Teter and Hemley first described these graphitic planes as a hexagonal organization of C3N3 triazine cycles. Because of its graphitic structure, the bulk modulus is only 51 GPa [22,26].
We propose m-C3N4 (m denotes Monoclinic symmetry, space group: Cm) and t-C3N4 (t denotes Tetragonal symmetry, space group: I-42m), whose structures are based on m-Si3N4 and t-Si3N4 [27], respectively, with C substituting for Si. The mechanical and electronic properties of m-C3N4 and t-C3N4 are presented in this work.

2. Computational Method

Density functional theory (DFT) [28,29] calculations within Vanderbilt ultrasoft pseudopotentials [30] were performed using the Cambridge Serial Total Energy Package (CASTEP) code [31]. For the exchange and correlation functional, we used the Perdew–Burke–Ernzerhof (PBE) version of the generalized gradient approximation (GGA) [32]. For α-C3N4, β-C3N4, d-ZB-C3N4, Pseudocubic-C3N4, Cubic-C3N4, g-C3N4, m-C3N4 and t-C3N4, an energy cutoff of 520 eV was used for the wave functions expansion. High dense k-point [33] sampling, with a grid spacing of less than 2π × 0.025 Å−1 (7 × 17 × 9 for m-C3N4, 11 × 11 × 6 for t-C3N4, 12 × 12 × 12 for d-ZB-C3N4, 12 × 12 × 12 for pseudocubic-C3N4, 8 × 8 × 8 for cubic-C3N4, 10 × 10 × 6 for g-C3N4, 7 × 7 × 8 for α-C3N4, and 7 × 7 × 18 for β-C3N4) in the Brillouin zone, was used. The equilibrium crystal structures were achieved by utilizing geometry optimization in the Broyden–Fletcher–Goldfarb–Shanno (BFGS) [34] minimization scheme. The self-consistent convergence of the total energy was 5 × 10−6 eV/atom; the maximum force on the atom was 0.01 eV/Å, the maximum ionic displacement was within 5 × 10−4 Å, and the maximum stress was within 0.02 GPa. The electronic properties of t-C3N4 and m-C3N4 were calculated using the Heyd–Scuseria–Ernzerhof (HSE06) [35,36] hybrid functional.

3. Results and Discussion

3.1. Structural Properties

The crystal structures of m-C3N4 and t-C3N4 are shown in Figure 1. There are 14 (six carbon atoms and eight nitrogen atoms) atoms in a conventional cell of m-C3N4 and t-C3N4. Within this structure of m-C3N4, three inequivalent carbon atoms occupy the (0.8392, 0.0, 0.5479), (0.2917, 0.0, 0.8418) and (0.9864, 0.0, 0.3048) positions, and four inequivalent nitrogen atoms occupy the (0.2050, 0.0, 0.3856), (0.8843, 0.0, 0.8186), (0.3600, 0.0, 0.1253) and (0.5344, 0.5, 0.5771) positions, while for t-C3N4, two inequivalent carbon atoms occupy the (0.5, 0.5, 0.0) and (0.5, 0.0, 0.75) positions, and nitrogen atoms occupy the (0.7330, 0.2670, 0.8705) position, respectively. The basic building block of m-C3N4 is the six-membered zigzag carbon-nitrogen rings, which can be clearly observed in Figure 1a; the twelve-membered zigzag carbon-nitrogen rings in the [010] direction in the structure of m-C3N4 are shown in Figure 1b, while six-membered zigzag carbon-nitrogen rings and eight-membered gauche carbon-nitrogen rings exist in t-C3N4. The equilibrium lattice parameters of m-C3N4, t-C3N4, d-ZB-C3N4 (space group: P-43m), cubic-C3N4 (space group: I-43d), pseudocubic-C3N4 (space group: P-42m), g-C3N4 (space group: P-6m2), α-C3N4 (space group: P31c) and β-C3N4 (space group: P63/m) at ambient pressure are listed in Table 1. The calculated parameters of α-C3N4 and β-C3N4 are in excellent agreement with previous theoretical results (see Table 1).
The calculated pressure–volume relationships of m-C3N4 and t-C3N4, together with diamond, c-BN, and other C3N4 allotropes, are shown in Figure 2. The highest incompressibility along the c-axis is due to m-C3N4 in the C3N4 allotropes, while along the c-axis, m-C3N4 yields the lowest incompressibility at pressures from 0 to 87 GPa; along the b-axis, m-C3N4 yields the lowest incompressibility at pressures from 87 to 100 GPa. For the crystal structure, pseudocubic-C3N4 has the greatest incompressibility in the C3N4 allotropes discussed above, while m-C3N4 has the weakest incompressibility. However, the incompressibility of t-C3N4 is greater than that of c-BN and the incompressibility of m-C3N4 is weaker than that of c-BN.

3.2. Elastic Properties and Hardness

In an effort to assess the thermodynamic stability of two novel C3N4 allotropes, enthalpy change curves with pressure for various structures were calculated, as presented in Figure 3. The dashed line represents the enthalpy of the summary of diamond and α-N2. It can be clearly seen that g-C3N4 has the lowest minimum value of enthalpy, which is in good agreement with previous reports and supports the reliability of our calculations [43]. Pseudocubic-C3N4 has the greatest minimum value of enthalpy. The minimum value of total energy per formula unit of t-C3N4 is slightly larger than that of g-C3N4, α-C3N4, m-C3N4, and β-C3N4 but much smaller than those of pseudocubic-C3N4 and cubic-C3N4, indicating that t-C3N4 and m-C3N4 should be thermodynamically metastable phases.
For monoclinic symmetry and tetragonal symmetry, there are different independent elastic constants. Monoclinic symmetry has thirty independent elastic constants (C11, C22, C33, C44, C55, C66, C12, C13, C23, C15, C25, C35 and C46), while tetragonal symmetry has six independent elastic constants (C11, C33, C44, C66, C12 and C13). The mechanical stability criteria of monoclinic symmetry are given by [44,45]:
C i i > 0 , i = 1 ~ 6.
[ C 11 + C 22 + C 33 + 2 ( C 12 + C 13 + C 23 ) ] > 0
( C 33 C 55 C 35 2 ) > 0
( C 44 C 66 C 46 2 ) > 0
( C 22 + C 33 2 C 23 ) > 0
[ C 22 ( C 33 C 55 C 35 2 ) + 2 C 23 C 25 C 35 C 23 2 C 55 C 25 2 C 33 ] > 0
Ω = 2 [ C 15 C 25 ( C 33 C 12 C 13 C 23 ) + C 15 C 35 ( C 22 C 13 C 12 C 33 ) + C 25 C 35 ( C 11 C 23 C 12 C 13 ) ] [ C 15 2 ( C 22 C 33 C 23 2 ) + C 25 2 ( C 11 C 33 C 13 2 ) + C 35 2 ( C 11 C 22 C 12 2 ) ] + C 55 g > 0
g = C 11 C 22 C 33 C 11 C 23 2 C 22 C 13 2 C 33 C 12 2 + 2 C 12 C 13 C 23
The criteria for the mechanical stability of tetragonal symmetry are given by [44]:
C i i > 0 , i = 1 , 3 , 4 , 6.
( C 11 C 12 ) > 0
( C 11 + C 33 2 C 13 ) > 0
[ 2 ( C 11 + C 12 ) + C 33 + 4 C 13 ] > 0
The calculated elastic constants of α-C3N4, β-C3N4, t-C3N4, d-ZB-C3N4, pseudocubic-C3N4, cubic-C3N4 and m-C3N4 are listed in Table 2. Elastic constants under high pressure were also studied. The elastic constants under ambient pressure and high pressure of t-C3N4 and m-C3N4 satisfied the mechanical stability criteria of monoclinic symmetry and tetragonal symmetry. Namely, t-C3N4 and m-C3N4 are mechanically stable. To confirm the stability of t-C3N4 and m-C3N4, their dynamical stabilities should also be studied under ambient pressure and high pressures. Thus, we calculated the phonon spectra for m-C3N4 and t-C3N4 at 0 and 100 GPa, as shown in Figure 4. No imaginary frequencies are observed throughout the whole Brillouin zone, signaling dynamically the stabilities of m-C3N4 and t-C3N4. The calculated elastic modulus of α-C3N4, β-C3N4, d-ZB-C3N4, pseudocubic-C3N4, cubic-C3N4, t-C3N4 and m-C3N4 are listed in Table 3. Bulk modulus B and shear modulus G were calculated by using the Voigt–Reuss–Hill approximation [46,47,48]. The Voigt and Reuss approximation of monoclinic symmetry is calculated using the following equations [44]:
B V = 1 9 [ C 11 + C 22 + C 33 + 2 ( C 12 + C 13 + C 23 ) ]
B R = Ω ( a + b + c + d + e + f ) 1
a = ( C 33 C 55 C 35 2 ) ( C 11 + C 22 2 C 12 )
b = ( C 23 C 55 C 25 C 35 ) ( 2 C 12 2 C 11 C 23 )
c = ( C 13 C 35 C 15 C 33 ) ( C 15 2 C 25 )
d = ( C 13 C 55 C 15 C 35 ) ( 2 C 12 + 2 C 23 C 13 2 C 22 )
e = ( C 13 C 25 C 15 C 23 ) ( C 25 C 15 )
f = C 11 ( C 22 C 55 C 25 2 ) C 12 ( C 12 C 55 C 15 C 25 ) + C 15 ( C 12 C 25 C 15 C 22 ) + C 25 ( C 23 C 35 C 25 C 33 )
G V = 1 15 [ C 11 + C 22 + C 33 + 3 ( C 44 + C 55 + C 66 ) ( C 12 + C 13 + C 23 ) ]
G R = 15 { 4 ( f + h + i + j + k + l ) Ω + 3 [ g Ω + ( C 44 + C 66 ) C 44 C 66 C 46 2 ] } 1
h = ( C 33 C 55 C 35 2 ) ( C 11 + C 22 + C 12 )
i = ( C 23 C 55 C 25 C 35 ) ( C 11 C 12 C 23 )
j = ( C 13 C 35 C 15 C 33 ) ( C 15 + C 25 )
k = ( C 13 C 55 C 15 C 35 ) ( C 22 C 23 C 12 C 13 )
l = ( C 13 C 25 C 15 C 23 ) ( C 15 C 25 ) .
The Voigt and Reuss approximation of tetragonal symmetry is calculated using the following equations [44]:
B V = 1 9 [ 4 C 13 + C 33 + 2 ( C 11 + C 12 ) ]
B R = ( C 11 + C 12 ) C 33 2 C 13 2 C 11 + C 12 + 2 C 33 4 C 13
G V = 1 30 [ C 11 + C 12 + 2 C 33 4 C 13 + 3 C 11 + 12 C 44 + 6 C 66 3 C 12 ) ]
G R = 15 { 18 B V ( C 11 + C 12 ) C 33 2 C 13 2 + [ 6 C 11 C 12 + 6 C 44 + 3 C 66 ] } 1 .
The Hill approximation of monoclinic symmetry and tetragonal symmetry is calculated using the following equation:
B H = B V + B R 2 , G H = G V + G R 2 .
Young’s modulus and Poisson’s ratio can be calculated using the following formulas, respectively: E = 9BHGH/(3BH + GH), v = (3BH − 2GH)/(6BH + 2GH) [49]. The relationships between elastic constants and pressures are shown in Figure 5a,b. Most of them increase with pressure, whereas C66 and C15 of m-C3N4 decrease with pressure. The dependence of the elastic constants on pressure of C22 of m-C3N4, i.e., dC22/dP = 6.97, means that C22 of m-C3N4 increases fastest among all elastic constants.
The dependence of bulk modulus, shear modulus and Young’s modulus on pressure of m-C3N4 and t-C3N4 is 3.49, 0.59, and 2.15 and 3.41, 1.59, and 4.35, respectively. Young’s modulus of t-C3N4 increases faster than other elastic modulus, while the increase in the shear modulus of m-C3N4 is the slowest. At ambient pressure, the bulk modulus of α-C3N4, β-C3N4, m-C3N4 and t-C3N4 are 387 GPa, 406 GPa, 327 GPa and 375 GPa, respectively. The calculated hardness of α-C3N4, β-C3N4, m-C3N4, t-C3N4, Cubic-C3N4, d-ZB-C3N4, Pseudocubic-C3N4 and c-BN are shown in Table 3. The bulk modulus of t-C3N4 is 375 GPa, which is slightly larger than that of c-BN, while the bulk modulus of m-C3N4 is slightly smaller than that of c-BN. The hardness of m-C3N4 is only 37 GPa, which is approximately half of that of α-C3N4, β-C3N4, d-ZB-C3N4, pseudocubic-C3N4 and t-C3N4. The reason for this phenomenon is that the mechanical properties of m-C3N4 are not excellent compared with the other C3N4 allotropes and the bulk modulus, shear modulus and Young’s modulus are all smaller than those of other C3N4 allotropes.
In materials science, ductility is a solid material’s ability to deform under tensile stress. If a material is brittle, when subjected to stress, it will break without significant deformation (strain). Additionally, these material properties are dependent on pressure. Pugh [56] proposed a simple relationship to judge the plastic properties of materials based on their elastic modulus, i.e., B/G. If the ratio B/G is larger than 1.75, a material exhibits the ductile property; otherwise, the material is brittle. Moreover, Poisson’s ratio v is another criterion for judging the plastic properties of materials [57]. A larger v value (v > 0.26) for a material indicates ductility, while a smaller v value (v < 0.26) usually denotes brittleness. At ambient pressure, the ratio B/G and v of α-C3N4, β-C3N4, d-ZB-C3N4, Cubic-C3N4, Pseudocubic-C3N4, m-C3N4 and t-C3N4 are as listed in Table 3. The ratio B/G and v of four C3N4 allotropes are all less than 1.75 and 0.26, respectively, which indicates that the four C3N4 allotropes are all brittle. t-C3N4 has the most brittleness, while β-C3N4 has the least brittleness. The pressure dependence of B/G and Poisson’s ratio v are shown in Figure 5c,d, respectively. In Figure 5c,d, the B/G and v of m-C3N4 and t-C3N4 increase with increasing pressure. m-C3N4 is found to change from brittle to ductile at 71 GPa, while t-C3N4 does not change from brittle to ductile in this pressure range.
The elastic anisotropy of a solid is closely related to the possibility of inducing microcracks in materials and can be expressed by the universal anisotropic index (AU) [58]. The universal anisotropic index is defined as AU = 5GV/GR + BV/BR-6. The calculated results of universal anisotropic index are also shown in Table 3. The universal anisotropic index of α-C3N4 is only 0.073, which is approximately one-third that of β-C3N4, approximately one-sixth that of t-C3N4, and approximately one-sixteenth that of m-C3N4. Namely, α-C3N4 and m-C3N4 exhibit the smallest and largest elastic anisotropy in AU, respectively. The pressure dependence of the universal anisotropic index is shown in Figure 5e. The universal anisotropic index of m-C3N4 increases faster than that of t-C3N4. The reason for this phenomenon is that the difference between the value of Voigt and Reuss approximations of shear modulus for m-C3N4 are greater than that of t-C3N4. At 0 GPa (100 GPa), the Voigt approximation values of the shear modulus for m-C3N4 and t-C3N4 are 293.26 GPa (403.14 GPa) and 360.75 GPa (534.66 GPa), respectively. The Reuss approximation values of the shear modulus for m-C3N4 and t-C3N4 are 254.44 GPa (262.45 GPa) and 340.05 GPa (484.94 GPa) at 0 GPa (100 GPa), respectively. The difference between the values of the Voigt and Reuss approximations of the shear modulus for m-C3N4 ranges from 38.8 to 140.7 GPa at 0 GPa and 100 GPa, respectively. Nevertheless, the difference between the value of the Voigt and Reuss approximations of the shear modulus for t-C3N4 only ranges from 20.7 to 49.7 GPa at 0 GPa and 100 GPa, respectively. Thus, the universal anisotropic index m-C3N4 increases faster than that of t-C3N4.
To analyze the anisotropy of m-C3N4 and t-C3N4 more systematically, we will investigate the anisotropy of m-C3N4 and t-C3N4 for Poisson’s ratio, the shear modulus and Young’s modulus by utilizing the ELAM codes [20,59]. The two-dimensional representations of Poisson’s ratio in the xy plane, xz plane and yz plane for m-C3N4 and t-C3N4 are shown in Figure 6. The blue, red and cyan lines represent Poisson’s ratio at 0, 50 and 100 GPa, while the solid line and dash-dot line represent the minimum and maximum values of Poisson’s ratio in the xy plane, xz plane and yz plane, respectively. From Figure 6, it is clear that the anisotropy of Poisson’s ratio for m-C3N4 and t-C3N4 increases with increasing pressure. The maximum value of Poisson’s ratio for m-C3N4 is 0.47, 0.61 and 0.76 at 0, 50 and 100 GPa, while the minimum value of Poisson’s ratio for m-C3N4 is 0.01; the maximum and minimum values of Poisson’s ratio are the same for m-C3N4 in the xy plane, xz plane and yz plane. The maximum value of Poisson’s ratio for t-C3N4 is 0.30, 0.38 and 0.44 at 0, 50 and 100 GPa, respectively, while the minimum value of Poisson’s ratio for t-C3N4 is 0.0; the maximum and minimum values of Poisson’s ratio are the same for t-C3N4 in the xy plane, xz plane and yz plane. The difference between the maximum and minimum values of Poisson’s ratio for t-C3N4 and m-C3N4 shows that m-C3N4 exhibits greater anisotropy with respect to Poisson’s ratio.
The 2D representations of the shear modulus in the xy plane, xz plane and yz plane for m-C3N4 and t-C3N4 are illustrated in Figure 7. For m-C3N4 in Figure 7a–c, the maximum value of the shear modulus occurs in the deviation from the x axis or y axis of approximately 45 degrees in the xy plane or yz plane, respectively. The maximum value of the shear modulus occurs in the deviation from the x axis or z axis of approximately 15 degrees in the xz plane. Moreover, the minimum value of the shear modulus occurs along the y axis in the xy plane and yz plane, respectively. The maximum and the minimum values of m-C3N4 are 163 GPa and 455 GPa at ambient pressure, respectively, and the ratio Gmax/Gmin = 2.79. At the same time, the maximum and minimum values of m-C3N4 are 153 GPa (110 GPa) and 611 GPa (738 GPa), respectively, at 50 GPa (100 GPa); the ratio Gmax/Gmin = 3.99 at p = 50 GPa, and the ratio Gmax/Gmin = 6.71 at p = 100 GPa. The anisotropy of m-C3N4 increases with increasing pressure. The average shear modulus of m-C3N4 is 266 GPa, 312 GPa and 328 GPa, respectively. For t-C3N4 in Figure 7d–f, the maximum value of the shear modulus for t-C3N4 in the xy plane, xz plane and yz plane appears along the coordinate axis, while the minimum value of the shear modulus for t-C3N4 in the xy plane, xz plane and yz plane appears in the deviation from the coordinate axis of approximately 45 degrees. The maximum and minimum values of the shear modulus for t-C3N4 are 245 GPa, 291 GPa, 323 GPa and 428 GPa, 559 GPa, 669 GPa at 0 GPa, 50 GPa, 100 GPa, respectively. The ratio Gmax/Gmin = 1.75 at p = 0 GPa, the ratio Gmax/Gmin = 1.92 at p = 50 GPa, and the ratio Gmax/Gmin = 2.07 at p = 100 GPa. It is clear that the ratio Gmax/Gmin for t-C3N4 is much smaller than that for m-C3N4. In other words, m-C3N4 exhibits greater anisotropy than t-C3N4. This agrees well with our previous prediction of anisotropy with respect to the universal anisotropic index and Poisson’s ratio.
As a valid method to describe the elastic anisotropic behavior of a crystal completely, the 3D surface constructions of the directional dependences of reciprocals of Young’s modulus are practically useful. The results are shown in Figure 8 for Young’s modulus. For an isotropic system, the 3D directional dependence would exhibit a spherical shape, while the deviation degree from the spherical shape reflects the content of anisotropy [60]. In Figure 8a,c, the 3D shape of Young’s modulus shows that m-C3N4 exhibits greater anisotropy than t-C3N4. As the pressure increases, the anisotropy of Young’s modulus for m-C3N4 and t-C3N4 increases, but m-C3N4 still exhibits greater anisotropy than t-C3N4. To analyze the anisotropy of m-C3N4 and t-C3N4 in detail, the 2D representations of Young’s modulus in the xy plane, xz plane and yz plane for m-C3N4 and t-C3N4 are depicted in Figure 9. From Figure 9, it is clear that m-C3N4 has a larger anisotropy and that the anisotropy will become larger with increasing pressure.
Young’s modulus of t-C3N4 has the same value in different planes, while that of m-C3N4 has different values in different planes. For example, at ambient pressure, the maximum and minimum values of Young’s modulus for t-C3N4 are 928 GPa and 612 GPa in the xy plane, xz plane and yz plane, while at p = 100 GPa, they are 1510 GPa and 864 GPa, respectively. However, the maximum value of Young’s modulus is 996 GPa in the xy plane and yz plane for m-C3N4, but in the xz plane, it is 995 GPa, and the minimum value is always 476 GPa. At 100 GPa, the difference reaches a larger degree; in the xy plane and yz plane, the maximum value of Young’s modulus for m-C3N4 is 1638 GPa, while the maximum value is 1634 GPa. This also proves that m-C3N4 has a larger anisotropy from the other side.

3.3. Electronic Structures

Band theory is one of the most stringent tests of the physics of semiconductors. For example, silicon, calcite and copper all contain similar densities of electrons, but they have different physical properties, all inexplicable without quantum mechanics [61]. Thus, it is necessary to understand the band structure and density of states. The band structures and density of states of m-C3N4 and m-C3N4 at different pressures are shown in Figure 10. The band structure calculations show that m-C3N4 is a quasi-direct band gap semiconductor, with a band gap of 4.52 eV (see Figure 10a), and at 100 GPa, m-C3N4 remains a quasi-direct band gap semiconductor, with a band gap of 5.68 eV. t-C3N4 has a quasi-direct band gap of 4.21 eV at (0.322 0.322 0.0) along the M−Γ direction and Γ point, while the direct band gap is 4.22 eV at the Γ point. At 100 GPa, t-C3N4 has a quasi-direct band gap of 4.79 eV at (0.322 0.322 0.0) along the M−Γ direction and Γ point, while the direct band gap is 4.81 eV at the Γ point. At 0 GPa, the valence band maximum (VBM) of m-C3N4 is located at the Z point, the energy of VBM near the Fermi level of m-C3N4 is 10.37 eV, and the energy of the Z point near the Fermi level is 10.39 eV; thus, m-C3N4 is a quasi-direct band gap semiconductor. At 100 GPa, the Fermi level of m-C3N4 increases to 14.04 eV, and the energy of the Z point near the Fermi level of m-C3N4 is 14.02 eV. The VBM of m-C3N4 is located at the point along the Z and Γ directions; its energy is 14.04 eV. The conduction band minimum (CBM) is at the Γ point for m-C3N4 at 0 and 100 GPa. The energy of CBM is 14.91 and 19.72 eV. At 0 GPa, the Fermi level of m-C3N4 is 10.39 eV, which is slightly smaller than that of t-C3N4 (10.52 eV). For t-C3N4, the CBM is at the Γ point for t-C3N4 at 0 and 100 GPa; the energy of CBM is 14.73 and 18.82 eV, respectively. The VBM of t-C3N4 is located at the point along the M and Γ directions; the energy is 10.52 and 14.03 eV, respectively. Moreover, the energy of the Γ point near the Fermi level of t-C3N4 is 10.51 eV and 14.01 eV at 0 and 100 GPa, respectively. Thus, t-C3N4 is a quasi-direct band gap semiconductor. Interestingly, the band gaps of m-C3N4 and t-C3N4 both increase with increasing pressure. At 100 GPa, m-C3N4 increases by 25.61%, and t-C3N4 increases by 13.66% compared with that at 0 GPa.

4. Conclusions

In conclusion, we have predicted two novel C3N4 allotropes, i.e., m-C3N4 and t-C3N4, with space groups Cm and I-42m, which are both mechanically and dynamically stable up to at least 100 GPa. The bulk modulus of t-C3N4 is 375 GPa, which is slightly larger than that of c-BN, while the bulk modulus of m-C3N4 is slightly smaller than that of c-BN. The hardness of t-C3N4 is larger than that of c-BN, thereby making it a superhard material with potential technological and industrial applications. The ratio B/G and v of the two novel C3N4 phases are both less than 1.75 and 0.26, respectively, which indicates that the two novel C3N4 allotropes are both brittle. The B/G and v of m-C3N4 and t-C3N4 increase with increasing pressure. m-C3N4 is found to change from being brittle to ductile at 71 GPa, while t-C3N4 does not change from being brittle to ductile in this pressure range. The elastic anisotropy calculations show that m-C3N4 and t-C3N4 both exhibit large anisotropy with respect to Poisson’s ratio, the shear modulus and Young’s modulus and universal anisotropic index. The band structure calculations show that m-C3N4 and t-C3N4 are a quasi-direct-band-gap semiconductor and a quasi-direct-band-gap semiconductor, respectively. Moreover, the band gaps of m-C3N4 and t-C3N4 continue to be a quasi-direct band-gap and quasi-direct band gap at 100 GPa, respectively. The band gaps of m-C3N4 and t-C3N4 are 4.522 and 4.210 eV, respectively, and these materials are both wide-band-gap semiconductors. Due to their quasi-direct band gaps, they are attractive for luminescent device applications.

Acknowledgments

This work was supported by the Natural Science Foundation of China (No. 61474089), the Natural Science Basic Research plan in Shaanxi Province of China (No. 2016JM1026), and the Open fund of key laboratory of complex electromagnetic environment science and technology, China Academy of Engineering Physics (No. 2015-0214. XY.K).

Author Contributions

Qingyang Fan and Qun Wei designed the project; Qingyang Fan, Changchun Chai and Qun Wei performed the calculations; Qingyang Fan, Qun Wei, and Yintang Yang determined the results; and Qingyang Fan and Changchun Chai wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hannay, J.B. Artificial Diamonds. Nature 1880, 22, 255–257. [Google Scholar] [CrossRef]
  2. Bridgman, P.W. An experimental contribution to the problem of diamond synthesis. J. Chem. Phys. 1947, 15, 92–98. [Google Scholar] [CrossRef]
  3. Cote, M.; Cohen, M.L. Carbon nitride compounds with 1:1 stoichiometry. Phys. Rev. B 1997, 55, 5684–5688. [Google Scholar]
  4. Betranhandy, E.; Matar, S.F. A model study for the breaking of cyanogen out of CNx within DFT. Diam. Relat. Mater. 2006, 15, 1609–1613. [Google Scholar] [CrossRef]
  5. Wang, J.T.; Chen, C.F.; Kawazoe, Y. Orthorhombic carbon allotrope of compressed graphite: Ab initio calculations. Phys. Rev. B 2012, 85, 033410. [Google Scholar] [CrossRef]
  6. Wang, J.T.; Chen, C.F.; Kawazoe, Y. Low-temperature phase transformation from graphite to sp3 orthorhombic carbon. Phys. Rev. Lett. 2011, 106, 075501. [Google Scholar] [CrossRef] [PubMed]
  7. Umemoto, K.; Wentzcovitch, R.M.; Saito, S.; Miyake, T. Body-centered tetragonal C4: A viable sp3 carbon allotrope. Phys. Rev. Lett. 2010, 104, 125504. [Google Scholar] [CrossRef] [PubMed]
  8. Li, D.; Bao, K.; Tian, F.B.; Zeng, Z.W.; He, Z.; Liu, B.B.; Cui, T. Lowest enthalpy polymorph of cold-compressed graphite phase. Phys. Chem. Chem. Phys. 2012, 14, 4347–4350. [Google Scholar] [CrossRef] [PubMed]
  9. Li, Q.K.; Sun, Y.; Li, Z.Y.; Zhou, Y. Lonsdaleite—A material stronger and stiffer than diamond. Scr. Mater. 2011, 65, 229–232. [Google Scholar]
  10. Zhou, R.L.; Zeng, X.C. Polymorphic phases of sp3-hybridized carbon under cold compression. J. Am. Chem. Soc. 2012, 134, 7530–7538. [Google Scholar] [CrossRef] [PubMed]
  11. Zhang, M.; Liu, H.Y.; Du, Y.H.; Zhang, X.X.; Wang, Y.C.; Li, Q. Orthorhombic C32: A novel superhard sp3 carbon allotrope. Phys. Chem. Chem. Phys. 2013, 15, 14120–14125. [Google Scholar] [CrossRef] [PubMed]
  12. He, C.Y.; Sun, L.Z.; Zhang, C.X.; Peng, X.Y.; Zhang, K.W.; Zhong, J.X. New superhard carbon phases between graphite and diamond. Solid State Commun. 2012, 152, 1560–1563. [Google Scholar] [CrossRef]
  13. Sheng, X.L.; Yan, Q.B.; Ye, F.; Zheng, Q.R.; Su, G. T-carbon: A novel carbon allotrope. Phys. Rev. Lett. 2011, 106, 155703. [Google Scholar] [CrossRef] [PubMed]
  14. Wei, Q.; Zhang, M.G.; Yan, H.Y.; Lin, Z.Z.; Zhu, X.M. Structural, electronic and mechanical properties of Imma-carbon. EPL 2014, 107, 27007. [Google Scholar] [CrossRef]
  15. Cheng, C.; Lv, Z.L.; Cheng, Y.; Chen, X.R.; Cai, L.C. A possible superhard orthorhombic carbon. Diam. Relat. Mat. 2014, 43, 49–54. [Google Scholar] [CrossRef]
  16. Wentorf, R.H., Jr. Cubic form of boron nitride. J. Chem. Phys. 1957, 26, 956. [Google Scholar] [CrossRef]
  17. Gueorguiev, G.K.; Neidhardt, J.; Stafstrom, S.; Hultman, L. First-principles calculations on the role of CN precursors for the formation of fullerene-like carbon nitride. Chem. Phys. Lett. 2005, 401, 288–295. [Google Scholar] [CrossRef]
  18. Gueorguiev, G.K.; Neidhardt, J.; Stafstrom, S.; Hultman, L. First-principles calculations on the curvature evolution and cross-linkage in carbon nitride. Chem. Phys. Lett. 2005, 410, 228–234. [Google Scholar] [CrossRef]
  19. Liu, A.Y.; Cohen, M.L. Prediction of new low compressibility solids. Science 1989, 245, 841–842. [Google Scholar] [CrossRef] [PubMed]
  20. Fan, Q.Y.; Wei, Q.; Chai, C.C.; Yan, H.Y.; Zhang, M.G.; Lin, Z.Z.; Zhang, Z.X.; Zhang, J.Q.; Zhang, D.Y. Structural, mechanical, and electronic properties of P3m1-BCN. J. Phys. Chem. Solids 2015, 79, 89–96. [Google Scholar] [CrossRef]
  21. Haeuseler, H. X-ray investigations in the system CdIn2S4-CdIn2Se4. J. Solid State Chem. 1979, 29, 121–123. [Google Scholar] [CrossRef]
  22. Liu, A.Y.; Wentzcovitch, R.M. Stability of carbon nitride solids. Phys. Rev. B 1994, 50, 10362. [Google Scholar] [CrossRef]
  23. Teter, D.M.; Hemley, R.J. Low-compressibility carbon nitrides. Science 1996, 271, 53–55. [Google Scholar] [CrossRef]
  24. Mo, S.D.; Ouyang, L.; Ching, W.Y.; Tanaka, I.; Koyama, Y.; Riedel, R. Interesting physical properties of the new spinel phase of Si3N4 and C3N4. Phys. Rev. Lett. 1999, 83, 5046–5049. [Google Scholar] [CrossRef]
  25. Kroll, P. Pathways to metastable nitride structures. J. Solid State Chem. 2003, 176, 530–537. [Google Scholar] [CrossRef]
  26. Manyali, G.S.; Warmbier, R.; Quandt, A.; Lowther, J.E. Ab initio study of elastic properties of super hard and graphitic structures of C3N4. Comput. Mater. Sci. 2013, 69, 299–303. [Google Scholar] [CrossRef]
  27. Cui, L.; Hu, M.; Wang, Q.Q.; Xu, B.; Yu, D.L.; Liu, Z.Y.; He, J.L. Prediction of novel hard phases of Si3N4: First-principles calculations. J. Solid State Chem. 2015, 228, 20–26. [Google Scholar] [CrossRef]
  28. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864. [Google Scholar] [CrossRef]
  29. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef]
  30. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892R. [Google Scholar] [CrossRef]
  31. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Kristallogr. 2005, 220, 567–570. [Google Scholar] [CrossRef] [Green Version]
  32. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  33. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  34. Pfrommer, B.G.; Côté, M.; Louie, S.G.; Cohen, M.L. Relaxation of crystals with the quasi-newton method. J. Comput. Phys. 1997, 131, 233–240. [Google Scholar] [CrossRef]
  35. Krukau, A.V.; Vydrov, O.A.; Izmaylov, A.F.; Scuseria, G.E. Influence of the exchange screening parameter on the performance of screened hybrid functionals. J. Chem. Phys. 2006, 125, 224106. [Google Scholar] [CrossRef] [PubMed]
  36. Fan, Q.Y.; Chai, C.C.; Wei, Q.; Yang, Y.T. Two novel silicon phases with direct band gaps. Phys. Chem. Chem. Phys. 2016, 18, 12905–12913. [Google Scholar] [CrossRef] [PubMed]
  37. Ruan, L.W.; Xu, G.S.; Chen, H.Y.; Yuan, Y.P.; Jiang, X.; Lu, Y.X.; Zhu, Y.J. The elastic behavior of dense C3N4 under high pressure: First-principles calculations. J. Phys. Chem. Solids 2014, 75, 1324–1333. [Google Scholar] [CrossRef]
  38. He, D.W.; Zhang, F.X.; Zhang, X.Y.; Qin, Z.C.; Zhang, M.; Liu, R.P.; Xu, Y.F.; Wang, W.K. Synthesis of carbon nitride crystals at high pressures and temperatures. J. Mater. Res. 1998, 13, 3458–3462. [Google Scholar] [CrossRef]
  39. Molina, B.; Sansores, L.E. Eelectronic structure of six phases of C3N4: A theoretical approach. Mod. Phys. Lett. B 1999, 13, 193. [Google Scholar] [CrossRef]
  40. Martin-Gil, J.; Martin-Gil, F.J.; Sarikaya, M.; Qian, M.; Jose-Yacaman, M.; Rubio, A. Evidence of a low compressibility carbon nitride with defect-zincblende structure. J. Appl. Phys. 1997, 81, 2555. [Google Scholar] [CrossRef]
  41. Tang, X.; Hao, J.; Lia, Y.W. A first-principles study of orthorhombic CN as a potential superhard material. Phys. Chem. Chem. Phys. 2015, 17, 27821–27825. [Google Scholar] [CrossRef] [PubMed]
  42. Goglio, G.; Foy, D.; Demazeau, G. State of Art and recent trends in bulk carbon nitrides synthesis. Mater. Sci. Eng. R Rep. 2008, 58, 195–227. [Google Scholar] [CrossRef]
  43. Xu, Y.; Gao, S.P. Band gap of C3N4 in the GW approximation. Int. J. Hydrog. Energy 2012, 37, 11072–11080. [Google Scholar] [CrossRef]
  44. Wu, Z.J.; Zhao, E.J.; Xiang, H.P.; Hao, X.F.; Liu, X.J.; Meng, J. Crystal structures and elastic properties of superhard IrN2 and IrN3 from first principles. Phys. Rev. B 2007, 76, 054115. [Google Scholar] [CrossRef]
  45. Fan, Q.Y.; Wei, Q.; Yan, H.Y.; Zhang, M.G.; Zhang, D.Y.; Zhang, J.Q. A New Potential Superhard Phase of OsN2. Acta Phys. Pol. A 2014, 126, 740. [Google Scholar] [CrossRef]
  46. Voigt, W. Lehrburch der Kristallphysik; Teubner, B.G., Ed.; Johnson Reprint Corp: Leipzig, Germany, 1928. [Google Scholar]
  47. Reuss, A. Berechnung der Fließgrenze von Mischkristallen auf Grund der Plastizitätsbedingung für Einkristalle. Z. Angew. Math. Mech. 1929, 9, 49–58. [Google Scholar] [CrossRef]
  48. Hill, R. The elastic behaviour of a crystalline aggregate. Proc. Phys. Soc. Lond. Sect. A 1952, 65, 349. [Google Scholar] [CrossRef]
  49. Fan, Q.Y.; Chai, C.C.; Wei, Q.; Yang, Y.T.; Qiao, L.P.; Zhao, Y.B.; Zhou, P.K.; Xing, M.J.; Zhang, J.Q.; Yao, R.H. Mechanical and electronic properties of Ca1-xMgxO alloys. Mater. Sci. Semicond. Process. 2015, 40, 676–684. [Google Scholar] [CrossRef]
  50. Grimsditch, M.; Zouboulis, E.S.; Polian, A. Elastic constants of boron nitride. J. Appl. Phys. 1994, 76, 832. [Google Scholar] [CrossRef]
  51. Levine, J.B.; Nguyen, S.L.; Rasool, H.I.; Wright, J.A.; Brown, S.E.; Kaner, R.B. Preparation and properties of metallic, superhard rhenium diboride crystals. J. Am. Chem. Soc. 2008, 130, 16953–16958. [Google Scholar] [CrossRef] [PubMed]
  52. Zhao, J.; Fan, C. First-principles study on hardness of five polymorphs of C3N4. Phys. B Condens. Matter 2008, 403, 1956–1959. [Google Scholar] [CrossRef]
  53. He, J.L.; Guo, L.C.; Guo, X.J.; Liu, R.P.; Tian, Y.J.; Wang, H.T.; Gao, C.X. Predicting hardness of dense C3N4 polymorphs. Appl. Phys. Lett. 2006, 88, 101906. [Google Scholar] [CrossRef]
  54. Fan, Q.Y.; Wei, Q.; Yan, H.Y.; Zhang, M.G.; Zhang, Z.X.; Zhang, J.Q.; Zhang, D.Y. Elastic and electronic properties of Pbca-BN: First-principles calculations. Comput. Mater. Sci. 2014, 85, 80–87. [Google Scholar] [CrossRef]
  55. Huang, Q.; Yu, D.; Zhao, Z.; Fu, S.; Xiong, M.; Wang, Q.; Gao, Y.; Luo, K.; He, J.L.; Tian, Y.J. First-principles study of O-BN: A sp3-bonding boron nitride allotrope. J. Appl. Phys. 2012, 112, 053518. [Google Scholar] [CrossRef]
  56. Pugh, S.F. XCII. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Lond. Edinb. Dublin Philos. Mag. J. Sci. Ser. 7 1954, 45, 823. [Google Scholar] [CrossRef]
  57. Duan, Y.H.; Sun, Y.; Peng, M.J.; Zhou, S.G. Anisotropic elastic properties of the Ca–Pb compounds. J. Alloy. Compd. 2014, 595, 14–21. [Google Scholar] [CrossRef]
  58. Ranganathan, S.I.; Ostoja-Starzewski, M. Universal elastic anisotropy index. Phys. Rev. Lett. 2008, 101, 055504. [Google Scholar] [CrossRef] [PubMed]
  59. Marmier, A.; Lethbridge, Z.A.D.; Walton, R.I.; Smith, C.W.; Parker, S.C.; Evans, K.E. ElAM: A computer program for the analysis and representation of anisotropic elastic properties. Comput. Phys. Commun. 2010, 181, 2102–2115. [Google Scholar] [CrossRef] [Green Version]
  60. Hu, W.C.; Liu, Y.; Li, D.J.; Zeng, X.Q.; Xu, C.S. First-principles study of structural and electronic properties of C14-type Laves phase Al2Zr and Al2Hf. Comput. Mater. Sci. 2014, 83, 27–34. [Google Scholar] [CrossRef]
  61. Singleton, J. Band Theory and Electronic Properties of Solids; Oxford University Press: New York, NY, USA, 2001. [Google Scholar]
Figure 1. The crystal structures of: m-C3N4 (a,b); and t-C3N4 (c) (black spheres denote carbon atoms, blue spheres denote nitrogen atoms).
Figure 1. The crystal structures of: m-C3N4 (a,b); and t-C3N4 (c) (black spheres denote carbon atoms, blue spheres denote nitrogen atoms).
Materials 09 00427 g001
Figure 2. The lattice constants a/a0, b/b0, c/c0 and V/V0 of compression as functions of pressure and temperature for: m-C3N4 (a); and t-C3N4 (b).
Figure 2. The lattice constants a/a0, b/b0, c/c0 and V/V0 of compression as functions of pressure and temperature for: m-C3N4 (a); and t-C3N4 (b).
Materials 09 00427 g002
Figure 3. Calculated enthalpies of different C3N4 structures relative to diamond and α-N2 as a function of pressure.
Figure 3. Calculated enthalpies of different C3N4 structures relative to diamond and α-N2 as a function of pressure.
Materials 09 00427 g003
Figure 4. Phonon spectra for: m-C3N4 (a,b); and t-C3N4 (c,d).
Figure 4. Phonon spectra for: m-C3N4 (a,b); and t-C3N4 (c,d).
Materials 09 00427 g004
Figure 5. Elastic constants of m-C3N4 (a) and t-C3N4 (b) as a function of pressure, and the ratio B/G (c); Poisson’s ratio v (d); and AU (e) of m-C3N4 and t-C3N4 as a function of pressure.
Figure 5. Elastic constants of m-C3N4 (a) and t-C3N4 (b) as a function of pressure, and the ratio B/G (c); Poisson’s ratio v (d); and AU (e) of m-C3N4 and t-C3N4 as a function of pressure.
Materials 09 00427 g005
Figure 6. 2D representation of Poisson’s ratio in the xy plane, xz plane and yz plane for: m-C3N4 (ac); and t-C3N4 (df). The solid line represents the minimum, and the dashed line represents the maximum. The blue, red and cyan lines represent Poisson’s ratio at p = 0, 50 and 100 GPa, respectively.
Figure 6. 2D representation of Poisson’s ratio in the xy plane, xz plane and yz plane for: m-C3N4 (ac); and t-C3N4 (df). The solid line represents the minimum, and the dashed line represents the maximum. The blue, red and cyan lines represent Poisson’s ratio at p = 0, 50 and 100 GPa, respectively.
Materials 09 00427 g006
Figure 7. 2D representation of shear modulus in the xy plane, xz plane and yz plane for: m-C3N4 (ac); and t-C3N4 (df).
Figure 7. 2D representation of shear modulus in the xy plane, xz plane and yz plane for: m-C3N4 (ac); and t-C3N4 (df).
Materials 09 00427 g007
Figure 8. The directional dependence of Young’s modulus for: m-C3N4 (a,b); and t-C3N4 (c,d).
Figure 8. The directional dependence of Young’s modulus for: m-C3N4 (a,b); and t-C3N4 (c,d).
Materials 09 00427 g008
Figure 9. 2D representation of Young’s modulus in the xy plane, xz plane and yz plane for: m-C3N4 (ac); and t-C3N4 (df). The blue, red and cyan lines represent Poisson’s ratio at p = 0, 50 and 100 GPa, respectively.
Figure 9. 2D representation of Young’s modulus in the xy plane, xz plane and yz plane for: m-C3N4 (ac); and t-C3N4 (df). The blue, red and cyan lines represent Poisson’s ratio at p = 0, 50 and 100 GPa, respectively.
Materials 09 00427 g009
Figure 10. Electronic band structure of: m-C3N4 (a,b); and t-C3N4 (c,d) with HSE06.
Figure 10. Electronic band structure of: m-C3N4 (a,b); and t-C3N4 (c,d) with HSE06.
Materials 09 00427 g010
Table 1. Calculated lattice parameters (Å) of m-C3N4, t-C3N4, d-ZB-C3N4, Cubic-C3N4, Pseudocubic-C3N4 and g-C3N4.
Table 1. Calculated lattice parameters (Å) of m-C3N4, t-C3N4, d-ZB-C3N4, Cubic-C3N4, Pseudocubic-C3N4 and g-C3N4.
MaterialsThis WorkOther Works
abca, b, c
α-C3N46.5124.742a = 6.489, c = 4.729 a; a = 6.425, c = 4.715 b; a = 6.467, c = 4.710 c;
a = 6.453, c = 4.699 d; a = 6.47, c = 4.71 e
β-C3N46.4492.422a = 6.426, c = 2.418 a; a = 6.419, c = 2.425 b; a = 6.402, c = 2.404 c;
a = 6.394, c = 2.397 d; a = 6.40, c = 2.40 e
d-ZB-C3N43.456a = 3.455 f; a = 3.52 e; a = 3.43 g
cubic-C3N45.398a = 5.395–5.444 h; a = 5.40 e
pseudocubic-C3N43.456a = 3.41–3.44 h
g-C3N44.7916.769a = 4.74, 6.72 e
m-C3N48.0322.4186.246
t-C3N43.4836.933
a Reference [37]; b Reference [38]; c Reference [23]; d Reference [39]; e Reference [40]; f Reference [41]; g Reference [22]; h Reference [42].
Table 2. The calculated elastic constants (GPa) of m-C3N4, t-C3N4, d-ZB-C3N4, cubic-C3N4, pseudocubic-C3N4 and cubic-BN.
Table 2. The calculated elastic constants (GPa) of m-C3N4, t-C3N4, d-ZB-C3N4, cubic-C3N4, pseudocubic-C3N4 and cubic-BN.
MaterialsC11C22C33C44C55C66C12C13C23C15C25C35C46
α-C3N4848906319335179131−2727
Reference [37]851906326334183129
β-C3N48521150286312228111
Reference [37]8331049289287259110
Reference [22]8341120305279138
d-ZB-C3N4791443184
Reference [41]794431184
Cubic-C3N4889518309
Reference [26]861469300
Pseudocubic-C3N4790792445444188187
Reference [37]804805439439183183
m-C3N456410028592093312135119536−5766048
t-C3N4702767428424212195
Cubic-BN823479185
Reference [50]-Exp.820480190
Table 3. The calculated elastic modulus (GPa), B/G, hardness Hv (GPa) and the universal anisotropic index of m-C3N4, t-C3N4, d-ZB-C3N4, cubic-C3N4, pseudocubic-C3N4 and cubic-BN.
Table 3. The calculated elastic modulus (GPa), B/G, hardness Hv (GPa) and the universal anisotropic index of m-C3N4, t-C3N4, d-ZB-C3N4, cubic-C3N4, pseudocubic-C3N4 and cubic-BN.
MaterialsBVBRGVGRBHGHB/GEvHvAU
α-C3N4386.83386.82338.65335.093873371.157840.16780.053
Reference [48]387.60387.60341.85338.963883401.147900.16820.043
β-C3N4406.11405.57330.77320.964063261.257720.18780.154
Reference [48]408.21407.84322.09311.124083171.297550.1963, 85 a0.176
d-ZB-C3N4386.59386.59387.23374.163873811.028610.13810.175
Reference [51]3873751.038500.1363
Cubic-C3N4502.48502.48426.82394.185024111.229690.18880.414
Reference [48]4964011.249480.1890, 87 b
Pseudocubic-C3N4388.21388.21387.66373.983883811.028610.13810.183
Reference [48]390.02390.02387.76376.883903821.028640.1370, 80 c0.144
m-C3N4332.22320.80293.26254.443272741.196430.17370.798
t-C3N4374.81374.47360.75340.053753501.078010.14800.305
c-BN397.43397.43414.90398.883974070.989100.12630.201
-400 d66 e, 64 f
a Reference [52]; b Reference [51]; c Reference [53]; d Reference [50]; e Reference [54]; f Reference [55].

Share and Cite

MDPI and ACS Style

Fan, Q.; Chai, C.; Wei, Q.; Yang, Y. Two Novel C3N4 Phases: Structural, Mechanical and Electronic Properties. Materials 2016, 9, 427. https://doi.org/10.3390/ma9060427

AMA Style

Fan Q, Chai C, Wei Q, Yang Y. Two Novel C3N4 Phases: Structural, Mechanical and Electronic Properties. Materials. 2016; 9(6):427. https://doi.org/10.3390/ma9060427

Chicago/Turabian Style

Fan, Qingyang, Changchun Chai, Qun Wei, and Yintang Yang. 2016. "Two Novel C3N4 Phases: Structural, Mechanical and Electronic Properties" Materials 9, no. 6: 427. https://doi.org/10.3390/ma9060427

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop