Next Article in Journal
Silk Materials Functionalized via Genetic Engineering for Biomedical Applications
Previous Article in Journal
Characterization and Bone Response of Carbonate-Containing Apatite-Coated Titanium Implants Using an Aqueous Spray Coating
Previous Article in Special Issue
A Review on Experimental Measurements for Understanding Efficiency Droop in InGaN-Based Light-Emitting Diodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two B-C-O Compounds: Structural, Mechanical Anisotropy and Electronic Properties under Pressure

1
Team of Micro & Nano Sensor Technology and Application in High-altitude Regions, Xizang Engineering Laboratory for Water Pollution Control and Ecological Remediation, School of Information Engineering, Xizang Minzu University, Xianyang 712082, China
2
School of Information Engineering, Chang’an University, Xi’an 710064, China
*
Author to whom correspondence should be addressed.
Materials 2017, 10(12), 1413; https://doi.org/10.3390/ma10121413
Submission received: 16 October 2017 / Revised: 16 November 2017 / Accepted: 8 December 2017 / Published: 11 December 2017
(This article belongs to the Special Issue Wide Bandgap Semiconductors: Growth, Properties and Applications)

Abstract

:
The structural, stability, mechanical, elastic anisotropy and electronic properties of two ternary light element compounds, B2CO2 and B6C2O5, are systematically investigated. The elastic constants and phonon calculations reveal that B2CO2 and B6C2O5 are both mechanically and dynamically stable at ambient pressure, and they can stably exist to a pressure of 20 GPa. Additionally, it is found that B2CO2 and B6C2O5 are wide-gap semiconductor materials with indirect energy gaps of 5.66 and 5.24 eV, respectively. The hardness calculations using the Lyakhov-Oganov model show that B2CO2 is a potential superhard material. Furthermore, the hardness of B6C2O5 is 29.6 GPa, which is relatively softer and more easily machinable compared to the B2CO2 (41.7 GPa). The elastic anisotropy results show that B6C2O5 exhibits a greater anisotropy in the shear modulus, while B2CO2 exhibits a greater anisotropy in Young’s modulus at ambient pressure.

1. Introduction

It has been found that the light elements such as B, C, N, and O can form strong covalently bonded materials that show intrinsic superhard characteristics. A series of new B-C-O compounds with multifunctional properties similar to B13C2 [1,2], B6O [2,3], B4C [3], BC [4,5], B2C [4,6], B3C [4,7], BC4 [4], B4C [4], B5C [4], B2CO [8,9], and B2C2O, B2C3O, and B2C5O [10], B4CO4 [11,12], 2D B-C-O alloys [13], BCxN [14,15,16,17,18,19], BCN [20,21,22], BN [20,23], BxO [24,25,26] and CxNy [27,28,29,30,31,32,33,34,35] has been designed by the theoretical method. These materials are almost all superhard materials.
Using the developed particle swarm optimization (PSO) algorithm for crystal structure prediction, Wang et al. [4] explored the possible crystal structures of B-C systems that are mechanically and dynamically stable. They found that with the exception of B4C and BC4, all boron carbides have high shear moduli (more than 240 GPa), indicating their strong resistance to shape change at constant volume. Theoretical Vickers hardness calculations showed that these boron carbides are potential superhard materials because the predicted hardnesses exceed 40 GPa. Zhang et al. [20] designed different kinds of superhard materials using the Crystal structure AnaLYsis by Particle Swarm Optimization (CALYPSO) algorithm. They found superhard phases in binary B-N compounds (such as Pct-BN, Pbca, Z-BN, BC8-BN, and M-BN) and ternary B-C-N compounds (such as I-4m2 BCN, Imm2 BCN [21], and P3m1 BCN [22]). Pbca-BN with an orthorhombic structure was investigated by first-principles calculations by Fan et al. [23], who found that Pbca-BN has the shear modulus of 316 GPa, bulk modulus of 344 GPa, large Debye temperature of 1734 K, and hardness of 60.1 GPa. Ternary B-C-N compounds I-4m2 BCN, Imm2 BCN, and P3m1 BCN were investigated by first-principles calculations by Fan et al. [21,22] and were all found to be potential superhard materials. Electronic structure studies showed that BCN materials in the P3m1 and I-4m2 phases are indirect semiconductors with band gaps of 4.10 eV and 0.45 eV, respectively, while the Imm2 phase is a direct semiconductor with a band gap of 2.54 eV.
Recently, Zhang et al. [10] performed extensive structure searches to explore the potential energetically stable B2CxO (x ≥ 2) phases at ambient pressure using the current developed CALYPSO algorithm. B2C2O, B2C3O, and B2C5O in the tetragonal I41/amd, I-4m2, and P-4m2 phases, respectively, were reported in ref. [10]. The phonon dispersion and formation enthalpy calculations revealed that B2C2O, B2C3O, and B2C5O are all dynamically stable and can be synthesized at ambient conditions. The hardnesses of B2C2O, B2C3O, and B2C5O are 57, 62, and 68 GPa using Gao’s model [36], and they are all potential superhard materials. B2CO, a potential superhard material in the B-C-O system, can stably exist in several different phases. tP4-B2CO, with the P-4m2 phase and tI16-B2CO with the I-42d phase have been reported by Li et al. [9], while oP8-B2CO with the Pmc21 phase was systematically investigated by Liu et al. [8] The elastic constants, phonon dispersion spectra and formation enthalpies confirmed the mechanical, dynamical and thermodynamic stabilities, respectively, of oP8-B2CO, tP4-B2CO, and tI16-B2CO. Calculation of the Vickers hardness of oP8-B2CO shows that oP8-B2CO is a superhard material, and the Vickers hardness of oP8-B2CO (47.70 GPa) is much closer to tP4-B2CO (49.24 GPa) and tI16-B2CO (49.56 GPa). The band structure calculations illustrate that all B2CO phases are semiconductor materials with indirect band gaps, and oP8-B2CO has the widest band gap (3.540 eV), greater than those of tP4-B2CO (1.658 eV) and tI16-B2CO (2.988 eV).
B4CO4, found in the I-4 space group, is a novel tetragonal thermodynamically stable phase, and two low-enthalpy metastable compounds (B6C2O5 in P1 phase, B2CO2 in C2/m phase) have been discovered by Wang et al. [12] using the widely used and evolutionary Universal Structure Predictor: Evolutionary Xtallography (USPEX) algorithm. The structural, stability, mechanical, elastic anisotropy and electronic properties of B4CO4 have been systematically investigated as described in ref. [11,12]. However, the structural, stability, mechanical, and elastic anisotropy properties of B2CO2 and B6C2O5 have not been reported. Therefore, in this work, we systematically investigate the structural, stability, mechanical, elastic anisotropy and electronic properties of B2CO2 and B6C2O5.

2. Theoretical Methods

The calculations were performed using density functional theory with the exchange-correlation functional treated using the generalized gradient approximation (GGA) in the Perdew-Burke-Ernzerhof (PBE) form [37] and the Perdew-Burke-Ernzerhof for solids (PBEsol) form [38], and with local density approximation (LDA) based on the data of Ceperley and Alder as parameterized by Perdew and Zunger (CA-PZ) [39,40]. The calculations in this work were performed using the Cambridge Serial Total Energy Package (CASTEP) code [41]. Structural optimizations were conducted using the Broyden-Fletcher-Goldfarb-Shanno (BFGS) minimization algorithm [42]. A high-density k-point [43] sampling with a grid spacing of less than 2π × 0.025 Å−1 (4 × 16 × 7 for B2CO2, 9 × 9 × 9 for B2CO2) in the Brillouin zone was used. The self-consistent convergence criterion for 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 HSE06 hybrid functional [44] was used for the calculations of electronic structures of B2CO2 and B6C2O5.

3. Results and Discussion

3.1. Structural Properties

The crystal structures of B2CO2 and B6C2O5 are shown in Figure 1a–d, respectively. For the structure shown in Figure 1a, there are two inequivalent oxygen atom positions (O1 in (0.1337, 0.0, 0.0474), red; O2 in (0.9735, 0.5, 0.2763), orange), two inequivalent boron atom positions (B1 in (0.8891, 0.0000, 0.2433), light green; B2 in (0.1298, 0.5000, 0.2027), dark green), and an equivalent carbon atom (C in (0.2497, 0.0000, 0.5855), blue) in B2CO2. In the structure shown in Figure 1c, each atom occupies a different position in B6C2O5 (B in (0.4165, 0.5981, 0.0121), (0.6796, 0.1566, 0.1255), (0.9807, 0.7214, 0.2895), (0.5464, 0.8782, 0.5674), (0.3288, 0.3515, 0.4784) and (0.1293, 0.0127, 0.8430); C in (0.4451, 0.9082, 0.8863) and (0.3015, 0.6221, 0.3228); and O in (0.5669, 0.1822, 0.4377), (0.8809, 0.8281, 0.5836), (0.7350, 0.4833, 0.0788), (0.9854, 0.0320, 0.1394), and (0.2105, 0.3326, 0.7576)). The calculated lattice parameters of B2CO2, B6C2O5, and other B-C-O compounds are listed in Table 1. The calculated lattice parameters of B2CO2 and B6C2O5 are in excellent agreement with the previous report [12], and the calculated lattice parameters of the other B-C-O compounds are also in excellent agreement with the previous report (see Table 1).
The relationships among the calculated lattice parameter ratios X/X0 (a/a0, b/b0, c/c0) and lattice volume ratios V/V0 of B2CO2 and B6C2O5 and pressure are shown in Figure 2, where a0, b0, c0 and V0 are the zero temperature and zero pressure equilibrium lattice constants and lattice volume, respectively. For B2CO2, it can be easily seen that the compression of the b-axis is the most difficult, whereas that of the a-axis is the easiest. For B6C2O5, similar to B2CO2, the b-axis is the most difficult to compress, while the c-axis is the easiest to compress. When the pressure increases, the compression along the b-axis of B6C2O5 is much larger than that along the b-axis of B2CO2. In addition, it is found that the lattice constants b/b0, c/c0 compression of B6C2O5 is changed at 10 GPa. This is because the lengths of some bonds decrease strongly along the lattice vector b-axis or the lattice vector c-axis. For example, there are three kinds of B-O bonds along the lattice vector c-axis; the bond lengths of B-O bonds are 1.571 Å (1.553 Å), 1.456 Å (1.459 Å) and 1.561 Å (1.543 Å) at 0 (5) GPa, respectively, and while the lengths of the first and the third bonds suddenly decreased to 1.516 Å and 1.508 Å under 10 GPa, the second bond length suddenly decreased to 1.489 Å under 10 GPa. Under 15 GPa, the first, second and third bond lengths dropped to 1.502 Å, 1.479 Å and 1.497 Å, respectively. From the above discussion, we can see that the abrupt change of the lattice constants c/c0 compression of B6C2O5 is due to the sudden decrease of the bond length of the B-O bonds along the lattice vector c-axis. The abrupt changes along the lattice vector b-axis is similar to that due to the abnormal change of the B-C bond length and the B-O bond. Their lattice constants changes show different trends with the pressure, which is related to the crystal structure and atomic composition. B6C2O5 exists in the P1 space group, whereas B2CO2 is found in the C2/m space group and the P1 space group has the lowest symmetry among the 230 space groups. In Figure 2b, for the lattice volume ratio, we predict that B2CO2 has better compressive resistance than B6C2O5. The results of Figure 2b confirm our prediction. Here, we can also predict that the bulk modulus of B2CO2 is greater than that of B6C2O5.

3.2. Stability

To demonstrate the dynamical stability of these compound, their phonon spectra are shown in Figure 3. At ambient pressure, there are no virtual frequencies in the entire Brillouin zone, proving that B2CO2 and B6C2O5 are both dynamically stable. When P = 20 GPa, there are still no virtual frequencies in the entire Brillouin zone; in other words, B2CO2 and B6C2O5 are still dynamically stable. The calculated elastic constants of B2CO2 and B6C2O5 under different pressures are listed in Table 2. The calculated elastic constants of B2CO2 and B6C2O5 at ambient pressure are in excellent agreement with the previously reported results [12]. There is no doubt that based on the data presented in Table 2, the elastic constants of B2CO2 and B6C2O5 satisfy the mechanical stability criteria [45], indicating that B2CO2 and B6C2O5 are mechanically stable. While dynamical and mechanical stabilities are verified by the phonon spectra and elastic constants, respectively, the formation enthalpy can be used to determine whether the new materials can be synthesized experimentally. Wang et al. have also described the possible synthetic routes of B2CO2 and B6C2O5 in Ref. [12]. With regard to elastic constants, compared to other B-C-O compounds, some elastic constants of B2CO2 and B6C2O5 are larger and some are smaller.

3.3. Mechanical Properties and Elastic Anisotropy

The primary elastic constants and elastic modulus values of B2CO2 and B6C2O5 as functions of pressure are shown in Figure 4, and the elastic modulus values for B2CO2, B6C2O5 and other B-C-O compounds are also listed in Table 2. Inspection of Figure 4 shows that almost all primary elastic constants and elastic moduli increase with increasing pressure. For B2CO2, the rate of increase of the elastic constants and elastic moduli remains almost constant, while for B6C2O5, the elastic constants and elastic moduli suddenly increase from 5 GPa to 10 GPa, and the rate of increase of elastic constants and elastic moduli remains almost constant under other pressures. As is well-known, the elastic constant represents the elastic properties of a material. Elastic constants provide a description of the relationship of stress and strain of different directions in an anisotropic medium. The elastic constants also obey Hooke’s law, and strain is proportional to stress, as expressed by Hooke’s law F = −kx, where k is a constant, called the stiffness coefficient. The stiffness coefficient is a complex physical quantity related to the material itself and the external conditions such as the temperature, so that the changes in the elastic constants for B6C2O5 observed in Figure 4c are also understandable. The lattice constant changes show different trends with the pressure, which are related to the differences in the crystal structure and atomic composition. B6C2O5 is found in the P1 space group, while B2CO2 is found in the C2/m space group, and the P1 space group has the lowest symmetry among the 230 space groups. For the P1 space group, there are 21 independent elastic constants, while for the C2/m space group, there are only 13 independent elastic constants. The hardnesses of B2CO2 and B6C2O5 were calculated by Wang et al. [11] using the Chen-Niu model [47] and the Lyakhov-Oganov model [48]. The results show that B2CO2 is a kind of superhard material. The Young’s modulus E is calculated as: E = 9BG/(3B + G). The calculated results for Young’s modulus for B2CO2 and B6C2O5 are also listed in Table 2. The Young’s moduli of B2CO2 and B6C2O5 increase with pressure, and the increase for B6C2O5 is greater than that for B2CO2. The increase for B6C2O5 was 23.88%, almost two times larger than that of B2CO2 (12.01%).
To analyze the anisotropy of B2CO2 and B6C2O5 more systematically, the anisotropies of the shear moduli and Young’s moduli of B2CO2 and B6C2O5 are investigated using the ELAM codes [22,45,49]. To better understand the mechanical and anisotropic properties of B2CO2 and B6C2O5, 3D surface figures of the shear moduli and Young’s moduli for B2CO2 and B6C2O5 are shown in Figure 5. The 3D surface figures represent the geometric figure that consists of the maximum or minimum value of the shear modulus or Young’s modulus in all directions of space. The magnitude of the shear modulus or Young’s modulus in all directions can be studied by using plane cutting, and the magnitude of the shear modulus or Young’s modulus in any plane can be represented by a two-dimensional graph. The 2D representations of the shear modulus and Young’s modulus are shown in Figure 6 and Figure 7, respectively. The 3D figure appears as a spherical shape for an isotropic material, while the deviation from the spherical shape is a measure of the degree of anisotropy [50]. It is clear that the shear and Young’s moduli of B2CO2 and B6C2O5 exhibit different degrees of anisotropy, and the anisotropies of the shear and Young’s moduli of B2CO2 and B6C2O5 increase with as the pressure increases from 0 GPa to 20 GPa. For example, a depression appears on the minimum value surface (green surface) of the shear modulus of B2CO2 at 20 GPa (see Figure 5b), but there is no dent under ambient pressure. In another example, at 20 GPa, a distinct dent appears on the three-dimensional Young’s modulus surface for B6C2O5 (see Figure 5h). Similarly, the indentation on the three-dimensional Young’s modulus surface for B6C2O5 is not so obvious at ambient pressure.
The sectional profiles are constructed on the basis of analysis of the geometrical characteristics of the 3D surfaces of the shear and Young’s moduli of B2CO2 and B6C2O5. Sectional drawings of the shear and Young’s moduli are shown in Figure 6 and Figure 7, respectively. The black, red and cyan lines represent the shear moduli at 0, 10 and 20 GPa, while the dash-dotted line and the solid line represent the maximum and minimum values of the shear modulus along different directions in the (001) plane (xy plane), (010) plane (xz plane), (100) plane (yz plane) and (111) plane, respectively. As shown in Figure 6, the anisotropy of B2CO2 and B6C2O5 in the shear modulus increases with increasing pressure. Examination of Figure 6a shows that along the direction of the x-axis, the maximum values of shear modulus shrink inwards at 20 GPa, but the maximum values of the shear modulus remain constant in the direction of the x-axis at ambient pressure. According to the geometrical shape of the profile, the anisotropy of B6C2O5 in the shear modulus is larger than that of B2CO2 at ambient pressure, while under high pressure, the anisotropy of B6C2O5 is smaller than that of B2CO2. This can be proved by comparing the ratio of the maximum to the minimum of the shear modulus (Gmax/Gmin). The maximum and the minimum values of the shear modulus and Young’s modulus for B2CO2 and B6C2O5 are listed in Table 3. The maximum value of shear modulus for B2CO2 increases with pressure, while the minimum value of the shear modulus for B2CO2 increases first and then decreases with pressure, so that the anisotropy of B2CO2 in the shear modulus becomes increasingly large. The anisotropy of B6C2O5 in the shear modulus decreases first and then increases.
The 2D representations of the Young’s moduli of B2CO2 and B6C2O5 in the (001) plane (xy plane), (010) plane (xz plane), (100) plane (yz plane) and (111) plane are displayed in Figure 7a–h, respectively. The anisotropy of Young’s modulus at each plane is also different, taking B2CO2 as an example. The Young’s modulus in the (010) plane (Emax/Emin = 628 GPa/414 GPa = 1.52) of B2CO2 exhibits the largest anisotropy compared to the other planes, while the (100) plane (Emax/Emin = 734 GPa/571 GPa = 1.29) exhibits the smallest anisotropy. This result can also be demonstrated by the ratio of the maximum to the minimum of the Young’s modulus at each plane. The maximum values of Young’s modulus for B2CO2 appear in the (001) plane and (100) plane, while the minimum value of Young’s modulus appears in the (010) plane. The maximum and minimum values appear in the same plane, whether at ambient pressure or at high pressure. Another interesting phenomenon is that the anisotropy of Young’s modulus in the (001) plane decreases with increasing pressure, and that of the (010) plane increases with increasing pressure, while those of the (100) plane and (111) plane decrease first and then increase. For the whole material, the anisotropies of the Young’s moduli of B2CO2 and B6C2O5 decrease first and then increase. In addition, although the maximum and minimum values appear in the same plane, whether at ambient pressure or at high pressure, the direction is changed. For B6C2O5, the maximum value appears at θ = 2.82, φ = 5.93 (the two angles are used to describe the unit vector, which is fully characterized by the angles θ (0, π), φ (0, 2π), as explained in more detail in Refs. [48,49]) at ambient pressure, where the angle is in radians. While the maximum value appears at θ = 1.59, φ = 2.90 at 10 GPa, for the pressure of 20 GPa, the maximum value appears at θ = 1.55, φ = 2.73. At these three pressures (0 GPa, 10 GPa, 20 GPa), the direction of the minimum appears to be different. However, for B2CO2, the maximum values all appear at θ = 1.57, φ = 4.73, whether at ambient pressure or at high pressure, but the direction of the minimum appears to have changed.

3.4. Electronic Properties

The electronic band structure is a significant physical property of a material that can be used to determine whether the material is a semiconductor, metal, or insulator. The electronic band structures of B2CO2 and B6C2O5 under different pressures are illustrated in Figure 8a–d, respectively. In Figure 8a–d, it is clear that B2CO2 and B6C2O5 are both indirect and wide band gap semiconductor materials with the band gaps of 5.63 and 5.24 eV, respectively. For B6C2O5, the conduction band minimum (CBM) and valence band maximum (VBM) are both located at the Dirac points in the Brillouin zone, the CBM of B2CO2 is also located at the Dirac point in the Brillouin zone, while the VBM of B2CO2 is not at the Dirac point in the Brillouin zone. The CBM and VBM are both located at the Q and F points in the Brillouin zone for B6C2O5, and the CBM of B2CO2 is also located at the V point in the Brillouin zone, while the VBM of B2CO2 is located at (−0.5000, −0.2143, 0.5000) along the L-M direction (L: (−0.5, 0.0, 0.5); M: (−0.5, −0.5, 0.5)) in the Brillouin zone.
The band gaps of B2CO2 and B6C2O5 under different pressures are shown in Figure 8e. The two materials show different trends for the change of the band gap with increasing pressure. The band gap of B6C2O5 increases with pressure, while the band gap of B2CO2 first increases and then decreases. The change trend of the band gap of B6C2O5 is similar to that of diamond and c-BN [22]: the band gap of B6C2O5, diamond and c-BN all increase with increasing pressure, but the changes for the band gaps of diamond and c-BN are relatively smooth, and the band gap changes for B6C2O5 are quite abrupt. Then, we investigated the origin of the pressure-driven abrupt increase of the band and the non-monotonic trend of first increase and then decrease of the band gap for B6C2O5 and B2CO2, respectively. We analyzed the trend in the Fermi level and the conduction band minimum with the change of pressure. The relationships between the Fermi level and the conduction band minimum for B2CO2 and B6C2O5 and the pressure are shown in Figure 8f. For the Fermi level and the conduction band minimum of B2CO2 and the conduction band minimum of B6C2O5, the changes are relatively smooth, but the energy at the CBM of B6C2O5 increases abruptly from 0 GPa to 5 GPa and from 5 GPa to 10 GPa. Therefore, the band gaps of B6C2O5 changes abruptly due to the sudden increase in the energy at the CBM leading to a sudden increase of the band gap from 5.309 eV under 5 GPa to 5.549 eV under 10 GPa.
To better understand why the band gap of B6C2O5 becomes larger with increased pressure, we also analyze the density of states of B6C2O5. The densities of states of B6C2O5 under different pressures are shown in Figure 9. Examination of Figure 9a shows that the valence band moves towards low energy, while the conduction band moves towards high energy in the 0–20 GPa range, so that the band gap of B6C2O5 increases with increased pressure. However, as shown in Figure 8e, the Fermi level increases with increased pressure, so, why does the band gap of B6C2O5 increase with the pressure in the 0–20 GPa range? According to Figure 9c, the density of states of the carbon atom almost does not vary with the pressure. Figure 9 b,d shows that, considering the density of states distribution of the oxygen atom, the peak value of the oxygen contribution to the conduction band increases with the increasing pressure, indicating that for the oxygen atoms under the action of external pressure, the local energy level is higher, resulting in reduced oxygen ions and leading to further changes of the band gap. Further consideration of the B atoms shows that the peak density of states increases with increasing pressure, decreasing the degree of localization of these atoms; however, the decrease in the density of states (from 1.07 state/eV/fu to 0.82 state/eV/fu when the pressure increases from 5 GPa to 10 GPa) is much smaller than the increase in the degree of the O atom localization (from 1.62 state/eV/fu to 3.24 state/eV/fu when the pressure increases from 5 GPa to 10 GPa), so that the results showed that the band gap of B6C2O5 becomes larger.

4. Conclusions

Based on first-principles calculations, we have systematically investigated two ternary light element compounds, B6C2O5 and B2CO2. We find that such two sp2 and sp3 hybridized structures could be obtained by other compounds and elements and that these structures are dynamically stable through calculations of the phonon spectra for B6C2O5 and B2CO2. The electronic band structures are also calculated, showing that B2CO2 and B6C2O5 are wide-gap semiconductor materials with indirect energy gaps of approximately 5.66 and 5.24 eV, respectively. The elastic anisotropy of B2CO2 and B6C2O5 phase has been demonstrated by the Young’s modulus and shear modulus along different crystal orientations. The elastic anisotropy results show that B6C2O5 exhibits a larger anisotropy in the shear modulus, while B2CO2 exhibits a larger anisotropy in Young’s modulus at ambient pressure. Another interesting phenomenon is that the anisotropy of the shear modulus of B2CO2 increases with increasing pressure, while the anisotropy of Young’s modulus of B2CO2, and the anisotropy of Young’s modulus and the shear modulus of B6C2O5 all first decrease and then increase with increasing pressure. The changes of the physical properties of B2CO2 and B6C2O5 show different trends with the pressure, and are strongly related to the crystal structure and atomic composition. The origin of band gap increases with increased pressure is explained from the viewpoint of the density of states. In addition, hardness calculations using the Lyakhov-Oganov model show that B2CO2 is a potential superhard material.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (No. 61762082), and Scientific Research Program of Shaanxi Education Department: Research and Design of Novel B-C-O Superhard Material. Wei Zhang (School of Microelectronics, Xidian University) is thanked for allowing to use the CASTEP code in Materials Studio.

Author Contributions

Liping Qiao designed the project; Liping Qiao and Zhao Jin performed the calculations; Liping Qiao and Zhao Jin determined the results; and Liping Qiao wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shirai, K.; Sakuma, K.; Uemura, N. Theoretical study of the structure of boron carbide B13C2. Phys. Rev. B 2014, 90, 064109. [Google Scholar] [CrossRef]
  2. Ektarawong, A.; Simak, S.I.; Hultman, L.; Birch, J.; Tasnádi, F.; Wang, F.; Alling, B. Effects of configurational disorder on the elastic properties of icosahedral boron-rich alloys based on B6O, B13C2, and B4C, and their mixing thermodynamics. J. Chem. Phys. 2016, 144, 134503. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, J.J.; Wang, Z.Y.; Jing, Y.Y.; Wang, S.Y.; Chou, C.F.; Hu, H.; Chiou, S.H.; Tsoo, C.C.; Su, W.S. Electronic structure and optical properties of boron suboxide B6O system: First-principles investigations. Solid State Commun. 2016, 244, 12–16. [Google Scholar] [CrossRef]
  4. Wang, D.Y.; Yan, Q.; Wang, B.; Wang, X.Y.; Yang, J.M.; Yang, G. Predicted boron-carbide compounds: A first-principles study. J. Chem. Phys. 2014, 140, 224704. [Google Scholar] [CrossRef] [PubMed]
  5. Fan, Q.Y.; Wei, Q.; Chai, C.C.; Yan, H.Y.; Zhang, M.G.; Zhang, Z.X.; Zhang, J.Q.; Zhang, D.Y. Structural, anisotropic and thermodynamic properties of boron carbide: First principles calculations. Indian J. Pure Appl. Phys. 2016, 54, 227–235. [Google Scholar]
  6. Fan, Q.Y.; Wei, Q.; Chai, C.C.; Yu, X.Y.; Liu, Y.; Zhou, P.K.; Yan, H.T.; Zhang, D.Y. First-principles study of structural, elastic, anisotropic, and thermodynamic properties of R3-B2C. Chin. J. Phys. 2015, 53, 100601. [Google Scholar] [CrossRef]
  7. Fan, Q.Y.; Wei, Q.; Chai, C.C.; Yang, Y.T.; Yu, X.H.; Liu, Y.; Zheng, J.P.; Zhou, P.K.; Zhang, D.Y. The elastic anisotropic and thermodynamic properties of I4mm-B3C. Acta Phys. Pol. A 2016, 129, 103–108. [Google Scholar] [CrossRef]
  8. Liu, C.; Zhao, Z.S.; Luo, K.; Hu, M.; Ma, M.D.; He, J.L. Superhard orthorhombic phase of B2CO compound. Diam. Relat. Mater. 2017, 73, 7–92. [Google Scholar] [CrossRef]
  9. Li, Y.W.; Li, Q.; Ma, Y.M. B2CO: A potential superhard material in the B-C-O system. EPL 2011, 95, 66006. [Google Scholar] [CrossRef]
  10. Zhang, M.G.; Yan, H.Y.; Zheng, B.B.; Wei, Q. Influences of carbon concentration on crystal structures and idealstrengths of B2CxO compounds in the B-C-O system. Sci. Rep. 2015, 5, 15481. [Google Scholar] [CrossRef] [PubMed]
  11. Zheng, B.B.; Zhang, M.G.; Wang, C.J. Exploring the mechanical anisotropy and ideal strengths of tetragonal B4CO4. Materials 2017, 10, 128. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, S.N.; Oganov, A.R.; Qian, G.R.; Zhu, Q.; Dong, H.F.; Dong, X.; Esfahani, M.M.D. Novel superhard B-C-O phases predicted from first principles. Phys. Chem. Chem. Phys. 2016, 18, 1859–1863. [Google Scholar] [CrossRef] [PubMed]
  13. Zhou, S.; Zhao, J.J. Two-dimensional B-C-O alloys: A promising 2D material for electronic devices. Nanoscale 2016, 8, 8910–8918. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, Y.; He, D.W.; Daemen, L.L.; Shen, R.B.; Schwarz, R.B.; Zhu, Y.; Bish, D.L.; Hung, J.; Zhang, J.; Shen, G.; et al. Superhard B-C-N materials synthesized in nanostructured bulks. J. Mater. Res. 2002, 17, 3139–3145. [Google Scholar] [CrossRef]
  15. He, J.L.; Tian, Y.J.; Yu, D.L.; Wang, T.S.; Liu, S.M.; Guo, L.C.; Li, D.C.; Jia, X.P.; Chen, L.X.; Zou, G.T.; et al. Orthorhombic B2CN crystal synthesized by high pressure and temperature. Chem. Phys. Lett. 2001, 340, 431–436. [Google Scholar] [CrossRef]
  16. Nakano, S.; Akaishi, M.; Sasaki, T.; Yamaoka, S. Segregative crystallization of several diamond-like phases from the graphitic BC2N without an additive at 7.7 GPa. Chem. Mater. 1994, 6, 2246–2251. [Google Scholar] [CrossRef]
  17. Knittle, E.; Kaner, R.B.; Jeanloz, R.; Cohen, M.L. High-pressure synthesis, characterization, and equation of state of cubic C-BN solid solutions. Phys. Rev. B 1995, 51, 12149. [Google Scholar] [CrossRef]
  18. Solozhenko, V.L.; Andrault, D.; Fiquet, G.; Mezouar, M.; Rubie, D.C. Synthesis of superhard cubic BC2N. Appl. Phys. Lett. 2001, 78, 1385–1387. [Google Scholar] [CrossRef]
  19. Komatsu, T.; Nomura, M.; Kakudate, Y.; Fujiwara, S. Synthesis and characterization of a shock-synthesized cubic B-C-N solid solution of composition BC2.5N. J. Mater. Chem. 1996, 6, 1799–1803. [Google Scholar] [CrossRef]
  20. Zhang, X.X.; Wang, Y.C.; Lv, J.; Zhu, C.Y.; Li, Q.; Zhang, M.; Li, Q.; Ma, Y.M. First-principles structural design of superhard materials. J. Chem. Phys. 2013, 138, 114101. [Google Scholar] [CrossRef] [PubMed]
  21. Fan, Q.Y.; Wei, Q.; Chai, C.C.; Zhang, M.G.; Yan, H.Y.; Zhang, Z.X.; Zhang, J.Q.; Zhang, D.Y. Elastic and electronic properties of Imm2- and I-4m2-BCN. Comput. Mater. Sci. 2015, 97, 6–13. [Google Scholar] [CrossRef]
  22. 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]
  23. 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]
  24. He, D.W.; Zhao, Y.S.; Daemen, L.; Qian, J.; Shen, T.D. Boron suboxide: As hard as cubic boron nitride. Appl. Phys. Lett. 2002, 81, 643–645. [Google Scholar] [CrossRef]
  25. Endo, T.; Sato, T.; Shimada, M. High-pressure synthesis of B2O with diamond-like structure. J. Mater. Sci. Lett. 1987, 6, 683–685. [Google Scholar] [CrossRef]
  26. Solozhenko, V.L.; Kurakevych, O.O.; Turkevich, V.Z.; Turkevich, D.V. Phase diagram of the B-B2O3 system at 5 GPa: Experimental and theoretical studies. J. Phys. Chem. B 2008, 112, 6683–6687. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. 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]
  29. Liu, A.Y.; Cohen, M.L. Prediction of new low compressibility solids. Science 1989, 245, 841–842. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, A.Y.; Wentzcovitch, R.M. Stability of carbon nitride solids. Phys. Rev. B 1994, 50, 10362. [Google Scholar] [CrossRef]
  31. Teter, D.M.; Hemley, R.J. Low-compressibility carbon nitrides. Science 1996, 271, 53–55. [Google Scholar] [CrossRef]
  32. 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]
  33. Molina, B.; Sansores, L.E. Eelectronic structure of six phases of C3N4: A theoretical approach. Mod. Phys. Lett. B 1999, 13, 193–201. [Google Scholar] [CrossRef]
  34. 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]
  35. Fan, Q.Y.; Chai, C.C.; Wei, Q.; Yang, Y.T. Two Novel C3N4 Phases: Structural, Mechanical and Electronic Properties. Materials 2016, 9, 427. [Google Scholar] [CrossRef] [PubMed]
  36. Gao, F.M.; He, J.L.; Wu, E.; Liu, S.M.; Yu, D.L.; Li, D.C.; Zhang, S.Y.; Tian, Y.J. Hardness of covalent crystals. Phys. Rev. Lett. 2003, 91, 015502. [Google Scholar] [CrossRef] [PubMed]
  37. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  38. Perdew, J.P.; Ruzsinszky, A.; Csonka, G.I.; Vydrov, O.A.; Scuseria, G.E.; Constantin, L.A.; Zhou, X.L.; Burke, K. Erratum: Restoring the density-gradient expansion for exchange in solids and surfaces. Phys. Rev. Lett. 2009, 102, 039902. [Google Scholar] [CrossRef]
  39. Ceperley, D.M.; Alder, B.J. Ground state of the electron gas by a stochastic method. Phys. Rev. Lett. 1980, 45, 566. [Google Scholar] [CrossRef]
  40. Perdew, J.P.; Zunger, A. Self-interaction correction to density-functional approximations for many-electron systems. Phys. Rev. B 1981, 23, 5048. [Google Scholar] [CrossRef]
  41. 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]
  42. 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]
  43. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  44. 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]
  45. Ma, Z.Y.; Han, Z.; Liu, X.H.; Yu, X.H.; Wang, D.T.; Tian, Y. Pnma-BN: Another boron nitride polymorph with interesting physical properties. Nanomaterials 2017, 7, 3. [Google Scholar] [CrossRef] [PubMed]
  46. Nye, J.F. Physical Properties of Crystals; Oxford University Press: Oxford, UK, 1985. [Google Scholar]
  47. Chen, X.Q.; Niu, H.Y.; Li, D.Z.; Li, Y.Y. Modeling hardness of polycrystalline materials and bulk metallic glasses. Intermetallics 2011, 19, 1275–1281. [Google Scholar] [CrossRef]
  48. Lyakhov, A.O.; Oganov, A.R. Evolutionary search for superhard materials: Methodology and applications to forms of carbon and TiO2. Phys. Rev. B 2011, 84, 092103. [Google Scholar] [CrossRef]
  49. 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]
  50. 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]
Figure 1. Lattice structure of B2CO2 (a); and along the [010] direction (b); and lattice structure of B6C2O5 (c); and along the [001] direction (d).
Figure 1. Lattice structure of B2CO2 (a); and along the [010] direction (b); and lattice structure of B6C2O5 (c); and along the [001] direction (d).
Materials 10 01413 g001
Figure 2. Lattice constants a/a0, b/b0, c/c0 compression as functions of pressure for B2CO2 and B6C2O5 (a); and primitive cell volume V/V0 for B2CO2 and B6C2O5 (b).
Figure 2. Lattice constants a/a0, b/b0, c/c0 compression as functions of pressure for B2CO2 and B6C2O5 (a); and primitive cell volume V/V0 for B2CO2 and B6C2O5 (b).
Materials 10 01413 g002
Figure 3. Phonon spectra of B2CO2 and B6C2O5 under different pressures: B2CO2 @ 0 GPa (a); B2CO2 @ 20 GPa (b); B6C2O5 @ 0 GPa (c); and B6C2O5 @ 20 GPa (d).
Figure 3. Phonon spectra of B2CO2 and B6C2O5 under different pressures: B2CO2 @ 0 GPa (a); B2CO2 @ 20 GPa (b); B6C2O5 @ 0 GPa (c); and B6C2O5 @ 20 GPa (d).
Materials 10 01413 g003
Figure 4. Elastic constants (a) and elastic modulus (b) as functions of pressure for B2CO2; and Elastic constants (c) and elastic modulus (d) as functions of pressure for B6C2O5.
Figure 4. Elastic constants (a) and elastic modulus (b) as functions of pressure for B2CO2; and Elastic constants (c) and elastic modulus (d) as functions of pressure for B6C2O5.
Materials 10 01413 g004
Figure 5. Directional dependence of the shear modulus at 0 GPa (a) and 20 GPa (b), and Young’s modulus at 0 GPa (c) and 20 GPa (d) of B2CO2; and Directional dependence of the shear modulus at 0 GPa (e) and 20 GPa (f), and Young’s modulus at 0 GPa (g) and 20 GPa (h) of B6C2O5.
Figure 5. Directional dependence of the shear modulus at 0 GPa (a) and 20 GPa (b), and Young’s modulus at 0 GPa (c) and 20 GPa (d) of B2CO2; and Directional dependence of the shear modulus at 0 GPa (e) and 20 GPa (f), and Young’s modulus at 0 GPa (g) and 20 GPa (h) of B6C2O5.
Materials 10 01413 g005
Figure 6. Planar projections of the directional dependence of the shear modulus in B2CO2, (a): (001) plane, (b): (010) plane, (c): (100) plane, and (d): (111) plane; and Planar projections of the directional dependence of the shear modulus in B6C2O5, (e): (001) plane, (f): (010) plane, (g): (100) plane, and (h): (111) plane.
Figure 6. Planar projections of the directional dependence of the shear modulus in B2CO2, (a): (001) plane, (b): (010) plane, (c): (100) plane, and (d): (111) plane; and Planar projections of the directional dependence of the shear modulus in B6C2O5, (e): (001) plane, (f): (010) plane, (g): (100) plane, and (h): (111) plane.
Materials 10 01413 g006aMaterials 10 01413 g006b
Figure 7. Planar projections of the directional dependence of the Young’s modulus in B2CO2, (a): (001) plane, (b): (010) plane, (c): (100) plane, and (d): (111) plane; and Planar projections of the directional dependence of the Young’s modulus in B6C2O5, (e): (001) plane, (f): (010) plane, (g): (100) plane, and (h): (111) plane.
Figure 7. Planar projections of the directional dependence of the Young’s modulus in B2CO2, (a): (001) plane, (b): (010) plane, (c): (100) plane, and (d): (111) plane; and Planar projections of the directional dependence of the Young’s modulus in B6C2O5, (e): (001) plane, (f): (010) plane, (g): (100) plane, and (h): (111) plane.
Materials 10 01413 g007
Figure 8. Band structures of: B2CO2 ((a): 0 GPa; and (b): 20 GPa); and B6C2O5 (c): 0 GPa; and (d): 20 GPa); band gaps as functions of pressure for B2CO2 and B6C2O5 (e); and Fermi level and CBM as functions of pressure for B2CO2 and B6C2O5 (f).
Figure 8. Band structures of: B2CO2 ((a): 0 GPa; and (b): 20 GPa); and B6C2O5 (c): 0 GPa; and (d): 20 GPa); band gaps as functions of pressure for B2CO2 and B6C2O5 (e); and Fermi level and CBM as functions of pressure for B2CO2 and B6C2O5 (f).
Materials 10 01413 g008
Figure 9. Density of states of B6C2O5 under different pressures: total density of states (a); partial density of states of B atom (b); partial density of states of C atom (c); and partial density of states of O atom (d).
Figure 9. Density of states of B6C2O5 under different pressures: total density of states (a); partial density of states of B atom (b); partial density of states of C atom (c); and partial density of states of O atom (d).
Materials 10 01413 g009
Table 1. Calculated lattice parameters of B2CO2, B6C2O5 and other B-C-O compounds.
Table 1. Calculated lattice parameters of B2CO2, B6C2O5 and other B-C-O compounds.
MaterialsSourcePBEPBEsolCA-PZ
a (Å)b (Å)c (Å)β (°)a (Å)b (Å)c (Å)β (°)a (Å)b (Å)c (Å)β (°)
B2CO2This work9.7732.4875.39590.99.7212.4795.36890.89.6062.4565.30890.9
ref. [12]9.7762.4895.39590.8
B6C2O5This work4.5034.5364.55798.14.4794.5264.50899.34.2784.4744.45799.3
ref. [12]4.5024.5384.55798.2
tP4-B2COThis work2.6562.6563.67890.02.6482.6483.65690.02.6182.6183.61990.0
ref. [9] 2.6232.6233.62390.0
tI16-B2COThis work3.7223.7227.49390.03.7023.7027.51590.03.6643.6647.37690.0
ref. [9] 3.6707.3947.39490.0
B2C2OThis work2.638 18.27490.02.627 18.17890.02.598 17.99690.0
ref. [10]2.647 18.27290.0
Table 2. Calculated elastic constants (GPa) and elastic modulus (GPa) of B2CO2 and B6C2O5 under different pressures (P: in GPa).
Table 2. Calculated elastic constants (GPa) and elastic modulus (GPa) of B2CO2 and B6C2O5 under different pressures (P: in GPa).
MaterialsPC11C22C33C44C55C66C12C13C23C15C25C35C46BGE
B2CO2058578461325621326411765136−771351−9287251583
562281364926722126912582145−811655−10306260608
1065884168626822327913296157−851557−10325267629
15694872718274226281141110168−901658−13344273648
20723902747260225283150126180−951364−18362272653
B6C2O5052851957120717021110240836−0.2−72−8227203469
55495475962161632151083593−10−1−87−10234207480
1064053463123319123713181122266−445274224528
1566557366524320424514291133318−497293234554
207046136972582102591551041423511−539312244581
tP4-B2CO0 1736 591240 25453157 311254
tI16-B2CO0 1600 646304 283182144 310265
B2C2O0 2763 590229 27415135 299264611
c-BN0 3779 446 165 370384
Pnma-BN0 439277067529927218799256116 298227543
Diamond0 31053 563 120 4315221116
1 Reference [9]; 2 Reference [10]; 3 Reference [22]; 4 Reference [46].
Table 3. Calculated the maximum values (in GPa) and minimum values (in GPa) of Young’s modulus and shear modulus and the Xmax/Xmin ratio for B2CO2 and B6C2O5.
Table 3. Calculated the maximum values (in GPa) and minimum values (in GPa) of Young’s modulus and shear modulus and the Xmax/Xmin ratio for B2CO2 and B6C2O5.
MaterialsP0 GPa10 GPa20 GPa
MaxMinRatioMaxMinRatioMaxMinRatio
B2CO2G3091701.823331771.883531752.02
E7344141.777854461.768374601.82
B6C2O5G2781401.962951611.833211701.88
E6033451.756574081.617174311.66

Share and Cite

MDPI and ACS Style

Qiao, L.; Jin, Z. Two B-C-O Compounds: Structural, Mechanical Anisotropy and Electronic Properties under Pressure. Materials 2017, 10, 1413. https://doi.org/10.3390/ma10121413

AMA Style

Qiao L, Jin Z. Two B-C-O Compounds: Structural, Mechanical Anisotropy and Electronic Properties under Pressure. Materials. 2017; 10(12):1413. https://doi.org/10.3390/ma10121413

Chicago/Turabian Style

Qiao, Liping, and Zhao Jin. 2017. "Two B-C-O Compounds: Structural, Mechanical Anisotropy and Electronic Properties under Pressure" Materials 10, no. 12: 1413. https://doi.org/10.3390/ma10121413

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