Next Article in Journal
Non-Equilibrium Plasma Processing for the Preparation of Antibacterial Surfaces
Next Article in Special Issue
Structural Transitions in Nanosized Zn0.97Al0.03O Powders under High Pressure Analyzed by in Situ Angle-Dispersive X-ray Diffraction
Previous Article in Journal
Two Independent Prospectively Planned Blinded Weibull Statistical Analyses of Flexural Strength Data of Zirconia Materials
Previous Article in Special Issue
A Stress-Induced Martensitic Transformation in Aged Ti49Ni51 Alloy after High-Velocity Impact
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanical Properties and Atomic Explanation of Plastic Deformation for Diamond-Like BC2

College of Physics and Optoelectronics Technology, Nonlinear Research Institute, Baoji University of Arts and Sciences, Baoji 721016, China
*
Author to whom correspondence should be addressed.
Materials 2016, 9(7), 514; https://doi.org/10.3390/ma9070514
Submission received: 28 April 2016 / Revised: 14 June 2016 / Accepted: 22 June 2016 / Published: 24 June 2016
(This article belongs to the Special Issue Computational Multiscale Modeling and Simulation in Materials Science)

Abstract

:
Motivated by a recently predicted structure of diamond-like BC2 with a high claimed hardness of 56 GPa (J. Phys. Chem. C 2010, 114, 22688–22690), we focus on whether this tetragonal BC2 (t-BC2) is superhard or not in spite of such an ultrahigh theoretical hardness. The mechanical properties of t-BC2 were thus further extended by using the first principles in the framework of density functional theory. Our results suggest that the Young’s and shear moduli of t-BC2 exhibit a high degree of anisotropy. For the weakest shear direction, t-BC2 undergoes an electronic instability and structural collapse upon a shear strain of about 0.11, with its theoretically ideal strength of only 36.2 GPa. Specifically, the plastic deformation under shear strain along the (110)[001] direction can be attributed to the breaking of d1 B–C bonds.

1. Introduction

Owing to the great demand from mechanical machining and semiconductor industries, the experimental and theoretical attempts to synthesizing and designing superhard materials have been performed intensively in recent decades. Although diamond is the hardest material with a Vickers hardness of 115 GPa along the (111) plane, the shortcoming of its poor resistance to oxidation as well as the reaction with ferrous metals restrict the applications of diamond. Those limitations have stimulated the continuous quest for novel superhard compounds with better thermal and chemical stabilities than pure diamond. In addition to possessing advanced electrical and optical properties as well as high hardness, the boron–carbon compounds have proven to be more highly resistant to oxygen and ferrous metals than similar carbon materials, which thus are unexceptionable substitutions for pure diamond. The typical B–C compound is boron-rich boron carbide (B4C), which is characterized by a unique combination of properties, prompting it as a choice of engineering material [1]. On the other hand, the boron-doped diamond lattice, which forms carbon-rich B–C material, changes it from an insulator into a p-type semiconductor with boron acting as a charge acceptor. Such fascinating electrical and mechanical properties has spurred researchers on to devote much effort to those carbon-rich B–C systems (BC2, BC3, BC5, and BC7) [2,3,4,5,6,7,8,9].
Recently, Zinin et al. synthesized a cubic BC3 phase with an sp3 bonding network at a pressure of 39 GPa and a temperature of 2200 K [4]. Solozhenko et al. had reported the synthesis of diamond-like BC5 with the highest boron content ever achieved under high pressure and high temperature condition [5]. The synthesized phase exhibits extreme Vickers hardness (71 GPa) and high thermal stability (up to 1900 K), which makes cubic BC5 an exceptional superabrasive overcoming diamond. However, the determinations of the synthesized B–C crystal structures are still open questions due to the similarity in both electronic and nuclear scattering cross sections for boron and carbon [1,10,11]. Theoretically, first-principle calculations combined with a state-of-the-art structure prediction have emerged as a powerful approach that complements experiments and achieved great success in the crystal structure prediction, especially in extreme physical conditions. Based on hypothetical configurations method and an ab initio evolutionary algorithm, Xu et al. [6] predicted a tetragonal lattice structure (space group I41/amd, No. 141) for BC2 (t-BC2) whose mechanical and dynamical stability have been confirmed by the criterions of elastic constants and phonon frequencies. Despite possessing the high theoretical Vickers hardness (56 GPa), we should carefully debate whether t-BC2 is superhard. The main reasons can be summarized as follows: (1) the absence of uniform and accurate hardness formula usually leads to the contradictory results for the same crystal for different models of hardness; and (2) the hardness is generally measured at finite strains where bonding characteristics of materials may change significantly [12], whereas the theoretical hardness estimated by the parameters at the equilibrium structure is thus not suitable to account for the mechanical strength of materials. Indeed, the ideal tensile and shear strengths of material evaluated when the lattice becomes unstable, which is the upper bound on the mechanical strength, are therefore more appropriate and stringent for the reflection of hardness than elastic parameters. In particular, the ideal shear strength describes the shear resistance of the system at the atomic level where plastic deformation occurs [13], so the ultimate hardness of a material may be assessed from its ideal shear strength and bonding nature, and this fact has been verified by a number of previous studies [14,15,16,17,18].
In the present paper, we perform the first-principles calculations to further investigate the structural and mechanical properties of the recently predicted t-BC2. The obtained orientation dependences of the Young’s and shear moduli are illustrated to show the mechanical anisotropy for t-BC2. The ideal tensile and shear strengths of t-BC2 are also estimated to provide a deeper insight into mechanical behavior and hardness.

2. Computational Methods

The total energy calculations were performed using density functional theory with the Perdew–Burke–Ernzerhof (PBE) exchange correlation in the framework of the generalized gradient approximation (GGA) as implemented in the Vienna ab initio simulation package (VASP) [19,20]. The electron and core interactions were described by the projector augmented-wave (PAW) method combined with the frozen core approximation [21], and the 2s22p1 and 2s22p2 were considered as valence electrons for B and C, respectively. The plane-wave basis set was truncated with an energy cutoff of 800 eV, and the Brillouin zone integration was generated using Monkhorst-Pack k point meshes [22] with a grid of 0.03 Å−1 and 0.025 Å−1 for total-energy and elastic constants calculations, respectively, which ensure that the enthalpy results were well converged to below 1 meV/f. u. The elastic constants were calculated by the strain–stress method, which has been successfully utilized previously [23,24]. The bulk modulus, shear modulus, Young’s modulus, and Poisson’s ratio were estimated via Voigt–Reuss–Hill approximation [25]. To obtain the strain–stress relationships and ideal strengths, t-BC2 cell was deformed continuously by increasing the displacement in the direction of the applied strain [15,26]. In addition, the 3D electron localization function distributions are illustrated by VESTA [27].

3. Results and Discussion

The crystal structure, as well as the dependences of the normalized lattice parameters and volume on pressure up to 100 GPa for t-BC2, is shown in Figure 1. The equilibrium lattice parameters for t-BC2 are a = 2.5227 Å and c = 11.9373 Å, which are in excellent agreement with the evaluated results reported in [6]. The optimized t-BC2 structure holds a tetragonal lattice with 12 atoms, and the two nonequivalent B and C atoms occupy the Wyckoff 4a (0.0, 0.0, 0.0) and 8e (0.0, 0.0, 0.34085) positions, respectively. The calculated bond lengths of C–C and B–C bonds are 1.502 Å and 1.663 Å, which are slightly different from the values (1.500 Å and 1.662 Å) proposed in [6], respectively. The estimated density is 3.06 g/cm3, which is lower than that of diamond (3.52 g/cm3). The remarkable consistency between the calculated results and [6] completely confirms the accuracy and reliability of the present calculations. The positive formation energy of t-BC2 (0.491 eV/atom), defined as E f = E tot ( BC 2 ) E ( solid B / atom ) 2 E ( diamond / atom ) , suggests that t-BC2 is meta-stable at ambient conditions. Compared with the BC3 structures synthesized experimentally, t-BC2 is more stable due to the greater formation energy of the BC3 (0.580 eV/atom). The calculated total energy of t-BC2 is −8.133 eV/atom, higher than that of the diamond-like BC3 (−8.351 eV/atom) and graphitic BC3 (−8.402 eV/atom) phases [28].
The estimated elastic constants, bulk modulus B, shear modulus G, and Young’s modulus E for t-BC2 are listed in Table 1 with the theoretical and experimental results of some previous B–C–N compounds for comparisons. For a stable tetragonal structure, the six independent elastic constants C11, C12, C13, C33, C44, and C66 should satisfy the necessary and sufficient Born–Huang elastic stability criteria as follows [37]:
C 11 > | C 12 | ,   2 C 13 2 < C 33 ( C 11 + C 12 ) , C 44 > 0 ,   C 66 > 0 .
Obviously, the calculated elastic constants of t-BC2 meet all the stability criteria, demonstrating the mechanical stability of t-BC2 at ambient pressure. Furthermore, the dependences of normalized lattice parameters on pressures up to 100 GPa for t-BC2, shown in Figure 1, indicate that t-BC2 structure along the c-axis is more incompressible than along the a-axis, which coincides with the result of elastic constants (C33 > C11). The compressibility of t-BC2 along the x- and z-directions as a function of pressure can be fitted by the least square method according to the lattice parameters and pressures, and the corresponding formulas are listed as follows:
a a 0 = 0.99958 1.00 × 10 3 P + 2.51 × 10 6 P 2 ;
c c 0 = 0.99941 7.14 × 10 4 P + 2.27 × 10 6 P 2 .
Typically, the hard material should possess a high bulk modulus to resist the volume deformation and a high shear modulus to support the shear deformation and thus enhance the resistance ability of material upon compression load. The calculated bulk modulus of t-BC2 is 333 GPa, much lower than that of diamond (432 GPa) but comparable to those of c-BN (376 GPa), d-BC3 (349 GPa), dl-BC3 (391 GPa), and dl-BC5 (376 GPa), suggesting that the t-BC2 phase is a greatly incompressible material. Physically, the shear modulus is more appropriate to evaluate the hardness of a material than the bulk modulus because the hardness tests measure plastic deformation of the material that appears to be closely linked to the deformation of a shear character [38]. Clearly, the shear modulus t-BC2 is 285 GPa, which is significantly smaller than that of superhard B–C–N compounds, such as diamond (517 GPa), c-BN (390 GPa), d-BC3 (318 GPa), dl-BC3 (344 GPa), and dl-BC5 (394 GPa). Compared with the diamond-like BCx phases of low pressure synthesis (i.e., dllp-B2C3 and dllp-BC4), the shear modulus of t-BC2 is much larger than those of the two phases (170 GPa and 19.3 GPa, respectively). Note that the bulk and shear moduli of B–C systems gradually decrease with the increasing concentrations of boron, which is consistent with the results of [30]. Therefore, we should carefully reappraise whether t-BC2 with large boron concentrations is superhard. According to Pugh’s criterion, the calculated result of G/B for t-BC2 is 0.847, larger than 0.571, indicating its brittle mechanical properties.
The mechanical anisotropy is the mechanical property of being directionally dependent, which can exert great influence on the properties of a physical mechanism. For a tetragonal structure, Young’s modulus for a tensile stress along an arbitrary [hkl] direction can be expressed as the following equation [39]:
E 1 = s 11 ( α 4 + β 4 ) + s 33 γ 4 + 2 s 12 α 2 β 2 + 2 s 13 ( β 2 γ 2 + α 2 γ 2 ) + s 44 ( β 2 γ 2 + α 2 γ 2 ) + s 66 α 2 β 2 ,
where α, β, and γ are the direction cosines of the tensile stress direction deduced from the transformed coordinate system with respect to the original coordinate system, and s11, s12, s13, s33, s44, and s66 are the independent elastic compliance constants given by Kelly et al. [39], which can be determined from the calculated elastic constants Cij. The shear modulus G on the (hkl) shear plane with shear stress applied along the [uvw] direction is given by
G 1 = 4 s 11 ( α 1 2 α 2 2 + β 1 2 β 2 2 ) + 4 s 33 γ 1 2 γ 2 2 + 8 s 12 α 1 α 2 β 1 β 2 + s 66 ( α 1 β 2 + α 2 β 1 ) 2 +   8 s 13 ( β 1 β 2 γ 1 γ 2 + α 1 α 2 γ 1 γ 2 ) + s 44 [ ( β 1 γ 2 + β 2 γ 1 ) 2 + ( α 1 γ 2 + α 2 γ 1 ) 2 ] ,
where (α1, β1, γ1) and (α2, β2, γ2) are the direction cosines of the [uvw] and [HKL] directions in the primitive coordinate system, respectively, and [HKL] directions denote the vector normal to the (hkl) shear plane. To gain deeper insight into the mechanical anisotropy of t-BC2, the three-dimension plots of the Young’s modulus as a function of the crystal orientation and its projections onto the (ab), (bc), and (ac) crystal planes are shown in Figure 2a,b, respectively. The distance between the origin of the coordinate and the surface or profile of the graph denotes the value of Young’s modulus along a certain direction. The large difference between the obtained shape and the sphere of a perfect isotropic crystal for the three-dimensional plots of Young’s modulus reveals that t-BC2 presents a high degree of anisotropy and hence is easily deformed along a certain direction under strain. Figure 2c,d illustrate the orientation dependence of Young’s modulus and shear modulus, respectively. From Figure 2c, we can clearly see that the maximum Young’s modulus (816 GPa) is 73.2% larger than the minimum Young’s modulus (471 GPa), further suggesting the high degree of anisotropy for t-BC2. The sequence of Young’s modulus along principle crystal orientations is summarized as follows: E[010] < E[001] < E[011] < E[111] < E[110]. As plotted in Figure 2d, the shear modulus along the (001) plane is independent of shear stress directions since the analytical result of shear modulus along this plane is described as G = 1/s44 = C44 = 395 GPa. The shear moduli along the (100) and ( 1 1 ¯ 0 ) planes decrease gradually with the increase in the angle of orientation, and the minimum shear modulus (199 GPa) distributes along the [110] direction within the ( 1 1 ¯ 0 ) basal plane.
Despite possessing ultrahigh elastic modulus, the feature of superhard for t-BC2 needs further confirmation, because the high values of elastic moduli can only represent high elastic stiffness—not a high degree of plastic harness [13]. Physically, the ideal strength, which is the ability to withstand an applied load up to yield, is more suitable to measure the hardness than the elastic modulus. The main reason is that the elastic modulus is a measure of elastic response that is non-permanent, and the material will return to its original shape when the applied load is removed, but the measurement of hardness is accompanied with plastic deformation involving the breaking of atomic bonds. To evaluate the strength determined by bond strength and the breaking nature under strain and essentially explore the atomistic origin of the structural deformation mechanisms, we calculated the ideal tensile and shear strengths of t-BC2 by means of applying a series of continuous strains along a specified direction, the calculated results are illustrated in Figure 3a,b. The ideal tensile strengths along the [001], [100], [110], and [111] directions are 139.9, 152.1, 70.1, and 41.6 GPa, respectively. Note that all the tensile strengths are larger than 40 GPa, and the minimum tensile strength occur in the body-diagonal [111] direction, which can be attributed to the absence of B–C and C–C bonds along the body-diagonal direction. Compared with the typical cubic BC3 (d-BC3) [31], the ideal tensile strength of t-BC2 along the [110] and [111] directions are slightly lower than those of the d-BC3[110] = 77.6 GPa, σ[111] = 52.5 GPa). Strikingly, the ideal tensile strength of t-BC2 along the [001] directions is even larger than that of the superhard d-BC3[001] = 107.6 GPa).
The shear stress responses in the (001), (100), and (110) planes were evaluated to search the easy cleavage plane of t-BC2. From Figure 3b, we can conclude that the ideal shear strengths show a high degree of anisotropy along the different shear directions, and the largest ideal shear strength in the ( 110 ) [ 1 1 ¯ 0 ] shear direction (98.6 GPa) is 2.7 times larger than the lowest ideal shear strength in the (110)[001] shear direction (36.2 GPa). The minimum of the ideal shear strength is 13.0% lower than the tensile strength along the [111] direction, suggesting that the failure mode in t-BC2 is dominated by the shear type. To obtain the fundamental mechanism of plastic deformation along the weakest shear direction, the bond lengths of d1 and d2 B–C bonds as a function of the shear strains in the (110)[001] direction are plotted in Figure 4. The development of the structure for t-BC2 under shear deformation along this direction and the local bonding structure are also sketched by the inset a, b, and c in Figure 4. Clearly, the d1 and d2 B–C bonds are identical at equilibrium (shear strain γ = 0). Then, with the increase of shear strain, the d1 B–C bond declines slightly but the d2 B–C bond rises rapidly. Once the shear strain goes beyond the critical strain (γ = 0.11279), the d2 bond increases to 2.061 Å abruptly, indicating the breaking of the d2 bond and the instability of the t-BC2 structure. In addition, the behavior of the covalent d2 bond breaking can be observed from the selected crystal structures before (inset a) and after (inset b) shear instability in Figure 4.
Note that the lowest ideal strength (36.2 GPa) is much lower than the calculated hardness (56.0 GPa [6]) estimated by Guo’s hardness formula [40]. Compared with typical boron oxide, the experimental Vickers’ hardness of superhard boron suboxide B6O (i.e., 38 GPa), harder than other boron oxides [41], is in excellent agreement with ab initio density functional theory of strain–stress curves for B6O (the lowest ideal shear strength of 38 GPa) [13,41,42], suggesting that the ideal shear strength is the most appropriate mechanical property for reflecting the hardness of materials. However, the calculated lowest ideal strength of t-BC2 shows a large discrepancy with the estimated hardness according to Guo’s hardness formula. The main reason for this discrepancy is that this hardness formula is correlated with bulk B and shear G moduli, which are estimated by the equilibrium structure, but the ideal shear strength focuses on the maximum shear stress at the atomic level where the crystal is far from equilibrium structure. To further clarify this disparity, the three-dimensional isosurfaces of the electron localization function (ELF) before and after the shear instability with an isovalue of 0.75 (a typical good number for characterization of covalent bondings) are plotted in Figure 5a,b, respectively. Apparently, the high electron localization appears between all adjacent B and C atoms when the shear strain is lower than the critical value of 0.11279. However, once the shear strain exceeds the critical value, the electron localization of B1–C1 and B2–C2 bonds vanish (denoted by red arrows), as illustrated in Figure 5b. The chemical bond defined by Bader is characterized by the bond critical point (BCP) whose charge density and Laplacian value reflect the type of chemical bond and the bond strength. Thus, we performed the Bader charge analysis, and the corresponding results are summarized in Table 2. Due to the ultrahigh bond strength, the C–C bond of t-BC2 exhibit large values of the electron density ρ and a relatively negative Laplacian value, 2 ρ , at the bond critical point, characteristic of primarily covalent bonding, which are comparable with those of diamond ( ρ = 1.60 −3 and 2 ρ = 15.24 −5 [43]), suggesting the high bond strength of the C–C bond in t-BC2 and hence hardly producing shear failure. However, the charge density at the d1 B–C bond critical point decrease significantly as the increase of shear strain, the Laplacian value of the d1 B–C bond critical point changes from −4.717 to −0.606, indicating the weakening covalent d1 bond under the shear strain and thus breaking beyond the critical shear strain. It can be concluded that the structural failure under shear deformation along the (110)[001] shear direction is rationalized by the breaking of the d1 bond. Furthermore, the minimum ideal strength of 36.2 GPa, lower than 40 GPa (the criterion of superhard material), indicates that the plastic deformation would take place along the crystal plane (110), leading to electronic instability and the structural collapse after a shear stress of 36.2 GPa.

4. Conclusions

In summary, we have investigated the mechanical anisotropy and strengths of t-BC2 by using first-principles calculations. The calculated Young’s and shear moduli of t-BC2 present a high degree of anisotropy and hence easily lead to structural failure. The shear strength for the slip system (110)[001] is the lowest among all the six-slip system, and this slip system leads to the breaking of the d1 B–C bonds in t-BC2, supported by the results of ELF and BCP calculations. These results demonstrate that t-BC2 tends to undergo an electronic instability with a shear stress lower than 40 GPa for the (110)[001] slip system. It should be stressed that the simplified models of hardness by either semi-empirical or ab initio methods are not appropriate for measuring the hardness of materials, because these theories usually introduce some parameters (such as bond energetics and electron density) that are based on the equilibrium structure, whereas plastic deformation in crystals occurs far from equilibrium upon bond breaking in the practical measurement of hardness. Thus, the ideal shear strength derived from the strain–stress curve, which is relevant for plastic deformation, is most appropriate for measuring the hardness of materials. The present work provides fundamental information for a better understanding of the structural stability and mechanical performance of this interesting material.

Acknowledgments

The work was supported partly by NSFC (Grant No. 11404008), the Natural Science Foundation of Shaanxi province of China (Grant No. 2016JM1016), the Education Committee Natural Science Foundation in Shaanxi Province of China (Grant No. 16JK1049), and the Baoji University of Arts and Sciences Key Research (Grant No. ZK16067).

Author Contributions

Baobing Zheng designed the project and performed the calculations; Baobing Zheng and Meiguang Zhang analyzed the results; Baobing Zheng, Meiguang Zhang, and Shaomei Chang wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Domnich, V.; Reynaud, S.; Haber, R.A.; Chhowalla, M. Boron Carbide: Structure, Properties, and Stability under Stress. J. Am. Ceram. Soc. 2011, 94, 3605–3628. [Google Scholar] [CrossRef]
  2. Ming, L.C.; Zinin, P.V.; Liu, X.R.; Nakamoto, Y.; Jia, R. Synthesis of dense BCx phases under high-pressure and high-temperature. J. Phys. Conf. Ser. 2010, 215, 012135. [Google Scholar] [CrossRef]
  3. Cermignani, W.; Paulson, T.E.; Onneby, C.; Pantano, C.G. Synthesis and characterization of boron-doped carbons. Carbon 1995, 33, 367–374. [Google Scholar] [CrossRef]
  4. Zinin, P.V.; Ming, L.C.; Ishii, H.A.; Jia, R.; Acosta, T.; Hellebrand, E. Phase transition in BCx system under high-pressure and high-temperature: Synthesis of cubic dense BC3 nanostructured phase. J. Appl. Phys. 2012, 111, 114905. [Google Scholar] [CrossRef]
  5. Solozhenko, V.L.; Kurakevych, O.O.; Andrault, D.; Le Godec, Y.; Mezouar, M. Ultimate Metastable Solubility of Boron in Diamond: Synthesis of Superhard Diamondlike BC5. Phys. Rev. Lett. 2009, 102, 015506. [Google Scholar] [CrossRef] [PubMed]
  6. Xu, L.; Zhao, Z.; Wang, L.-M.; Xu, B.; He, J.; Liu, Z.; Tian, Y. Prediction of a Three-Dimensional Conductive Superhard Material: Diamond-like BC2. J. Phys. Chem. C 2010, 114, 22688–22690. [Google Scholar] [CrossRef]
  7. Liu, H.; Li, Q.; Zhu, L.; Ma, Y. Superhard and superconductive polymorphs of diamond-like BC3. Phys. Lett. A 2011, 375, 771–774. [Google Scholar] [CrossRef]
  8. Li, Q.; Wang, H.; Tian, Y.; Xia, Y.; Cui, T.; He, J.; Ma, Y.; Zou, G. Superhard and superconducting structures of BC5. J. Appl. Phys. 2010, 108, 023507. [Google Scholar] [CrossRef]
  9. Liu, H.; Li, Q.; Zhu, L.; Ma, Y. Superhard polymorphs of diamond-like BC7. Solid State Commun. 2011, 151, 716–719. [Google Scholar] [CrossRef]
  10. Morosin, B.; Kwei, G.H.; Lawson, A.C.; Aselage, T.L.; Emin, D. Neutron powder diffraction refinement of boron carbides nature of intericosahedral chains. J. Alloy. Compd. 1995, 226, 121–125. [Google Scholar] [CrossRef]
  11. Jiménez, I.; Sutherland, D.G.J.; van Buuren, T.; Carlisle, J.A.; Terminello, L.J.; Himpsel, F.J. Photoemission and X-ray-absorption study of boron carbide and its surface thermal stability. Phys. Rev. B 1998, 57, 13167–13174. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Sun, H.; Chen, C. Superhard Cubic BC2N Compared to Diamond. Phys. Rev. Lett. 2004, 93, 195504. [Google Scholar] [CrossRef] [PubMed]
  13. Veprek, S.; Zhang, R.F.; Argon, A.S. Mechanical properties and hardness of boron and boron-rich solids. J. Superhard Mater. 2011, 33, 409–420. [Google Scholar] [CrossRef]
  14. Zhang, Y.; Sun, H.; Chen, C. Atomistic Deformation Modes in Strong Covalent Solids. Phys. Rev. Lett. 2005, 94, 145505. [Google Scholar] [CrossRef] [PubMed]
  15. Roundy, D.; Krenn, C.R.; Cohen, M.L.; Morris, J.W. Ideal Shear Strengths of fcc Aluminum and Copper. Phys. Rev. Lett. 1999, 82, 2713–2716. [Google Scholar] [CrossRef]
  16. Zhang, R.F.; Legut, D.; Niewa, R.; Argon, A.S.; Veprek, S. Shear-induced structural transformation and plasticity in ultraincompressible ReB2 limit its hardness. Phys. Rev. B 2010, 82, 104104. [Google Scholar] [CrossRef]
  17. Zheng, B.; Zhang, M.; Luo, H.-G. Mechanical anisotropy and origin of shear plastic deformation of tetragonal B4C4. Europhys. Lett. 2014, 108, 16001. [Google Scholar] [CrossRef]
  18. Zhang, M.; Yan, H.; Zhao, Y.; Wei, Q. Mechanical properties and atomistic deformation mechanism of spinel-type BeP2N4. Comput. Mater. Sci. 2014, 83, 457–462. [Google Scholar] [CrossRef]
  19. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  20. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  21. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  22. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  23. Milman, V.; Warren, M.C. Elasticity of hexagonal BeO. J. Phys. Condens. Matter 2001, 13, 241. [Google Scholar] [CrossRef]
  24. Zhang, M.; Yan, H.; Zheng, B.; Wei, Q. Influences of carbon concentration on crystal structures and ideal strengths of B2CxO compounds in the B-C-O system. Sci. Rep. 2015, 5, 15481. [Google Scholar] [CrossRef] [PubMed]
  25. Hill, R. The Elastic Behaviour of a Crystalline Aggregate. Proc. Phys. Soc. Lond. Sect. A 1952, 65, 349. [Google Scholar] [CrossRef]
  26. Zhang, R.F.; Sheng, S.H.; Veprek, S. First principles studies of ideal strength and bonding nature of AlN polymorphs in comparison to TiN. Appl. Phys. Lett. 2007, 91, 031906. [Google Scholar] [CrossRef]
  27. Momma, K.; Izumi, F. VESTA3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  28. Lowther, J.E. The Role Played by Computation in Understanding Hard Materials. Materials 2011, 4, 1104–1116. [Google Scholar] [CrossRef]
  29. Lowther, J.E.; Zinin, P.V.; Ming, L.C. Vibrational energies of graphene and hexagonal structured planar B–C complexes. Phys. Rev. B 2009, 79, 033401. [Google Scholar] [CrossRef]
  30. Nkambule, S.M.; Lowther, J.E. Crystalline and random “diamond-like” boron–carbon structures. Solid State Commun. 2010, 150, 133–136. [Google Scholar] [CrossRef]
  31. Zhang, M.; Liu, H.; Li, Q.; Gao, B.; Wang, Y.; Li, H.; Chen, C.; Ma, Y. Superhard BC3 in Cubic Diamond Structure. Phys. Rev. Lett. 2015, 114, 015502. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Y.-J.; Wang, C.-Y. Mechanical properties and electronic structure of superhard diamondlike BC5: A first-principles study. J. Appl. Phys. 2009, 106, 043513. [Google Scholar] [CrossRef]
  33. Dodd, S.P.; Saunders, G.A.; James, B. Temperature and pressure dependences of the elastic properties of ceramic boron carbide (B4C). J. Mater. Sci. 2002, 37, 2731–2736. [Google Scholar] [CrossRef]
  34. Lee, S.; Bylander, D.M.; Kleinman, L. Elastic moduli of B12 and its compounds. Phys. Rev. B 1992, 45, 3245–3247. [Google Scholar] [CrossRef]
  35. Zhang, R.F.; Veprek, S.; Argon, A.S. Anisotropic ideal strengths and chemical bonding of wurtzite BN in comparison to zincblende BN. Phys. Rev. B 2008, 77, 172103. [Google Scholar] [CrossRef]
  36. Zhang, R.F.; Lin, Z.J.; Veprek, S. Anisotropic ideal strengths of superhard monoclinic and tetragonal carbon and their electronic origin. Phys. Rev. B 2011, 83, 155452. [Google Scholar] [CrossRef]
  37. Mouhat, F.; Coudert, F.-X. Necessary and sufficient elastic stability conditions in various crystal systems. Phys. Rev. B 2014, 90, 224104. [Google Scholar] [CrossRef]
  38. Teter, D.M. Computational Alchemy: The Search for New Superhard Materials. MRS Bull. 1998, 23, 22–27. [Google Scholar] [CrossRef]
  39. He, Y.; Schwarz, R.B.; Migliori, A.; Whang, S.H. Elastic constants of single crystal γ-TiAl. J. Mater. Sci. 1995, 10, 1187–1195. [Google Scholar] [CrossRef]
  40. Guo, X.; Li, L.; Liu, Z.; Yu, D.; He, J.; Liu, R.; Xu, B.; Tian, Y.; Wang, H.-T. Hardness of covalent compounds: Roles of metallic component and d valence electrons. J. Appl. Phys. 2008, 104, 023503. [Google Scholar] [CrossRef]
  41. Kurakevych, O.O.; Solozhenko, V.L. Experimental study and critical review of structural, thermodynamic and mechanical properties of superhard refractory boron suboxide B6O. J. Superhard Mater. 2011, 33, 421–428. [Google Scholar] [CrossRef]
  42. Mukhanov, V.A.; Kurakevych, O.O.; Solozhenko, V.L. Thermodynamic model of hardness: Particular case of boron-rich solids. J. Superhard Mater. 2010, 32, 167–176. [Google Scholar] [CrossRef]
  43. Yu, R.; Zhan, Q.; Zhang, X.F. Elastic stability and electronic structure of pyrite type PtN2: A hard semiconductor. Appl. Phys. Lett. 2006, 88, 051913. [Google Scholar] [CrossRef]
Figure 1. The crystal structure and the dependence of normalized lattice constants and cell volume on pressure for t-BC2. The blue and black spheres denote B and C atoms, respectively.
Figure 1. The crystal structure and the dependence of normalized lattice constants and cell volume on pressure for t-BC2. The blue and black spheres denote B and C atoms, respectively.
Materials 09 00514 g001
Figure 2. (a,c) Orientation dependence of Young’s modulus E and (b) the corresponding projection in ab, ac, and bc planes; (d) for t-BC2,orientation dependence of the shear modulus of t-BC2.
Figure 2. (a,c) Orientation dependence of Young’s modulus E and (b) the corresponding projection in ab, ac, and bc planes; (d) for t-BC2,orientation dependence of the shear modulus of t-BC2.
Materials 09 00514 g002
Figure 3. Calculated strain–stress relations for t-BC2 in various tensile (a) and shear (b) directions.
Figure 3. Calculated strain–stress relations for t-BC2 in various tensile (a) and shear (b) directions.
Materials 09 00514 g003
Figure 4. Calculated bond lengths as a function of strain for t-BC2 under shear deformation along the (110)[001] directions. Insets: crystal structures before (a) and after (b) shear instability, and basic building blocks in t-BC2 (c). The dashed line represents the shear-induced structural deformation’s first occurrence.
Figure 4. Calculated bond lengths as a function of strain for t-BC2 under shear deformation along the (110)[001] directions. Insets: crystal structures before (a) and after (b) shear instability, and basic building blocks in t-BC2 (c). The dashed line represents the shear-induced structural deformation’s first occurrence.
Materials 09 00514 g004
Figure 5. Developments of ELF for t-BC2 during shear in the (110)[001] slip before (a) and after (b) shear instability.
Figure 5. Developments of ELF for t-BC2 during shear in the (110)[001] slip before (a) and after (b) shear instability.
Materials 09 00514 g005
Table 1. Calculated elastic constants Cij, bulk modulus B, shear modulus G, and Young’s modulus E (in units of GPa). Also shown is G/B ratio.
Table 1. Calculated elastic constants Cij, bulk modulus B, shear modulus G, and Young’s modulus E (in units of GPa). Also shown is G/B ratio.
CompoundsSourceC11C12C13C33C44C66BGEG/B
t-BC2Present5711732266123953243332826590.847
B4C4Theory 16561911675623113823242856600.879
dl-BC3Theory 2720206220788464268391344798
d-BC3Theory 3658195 393 349318731
dl-BC5Theory 4818156 442 376394876
B4CExperiment 5 240193456
Theory 656212470518 234
c-BNTheory 7786172 445 376390
DiamondTheory 81052122 555 432517
1 [17]; 2 [29,30]; 3 [31]; 4 [32]; 5 [33]; 6 [34]; 7 [35]; 8 [32,36].
Table 2. Bond critical point data of t-BC2 structure before and after shear instability. L is the bond length. ρ ( r C P ) and 2 ρ ( r CP ) are the charge density and its Laplacian at the corresponding critical points.
Table 2. Bond critical point data of t-BC2 structure before and after shear instability. L is the bond length. ρ ( r C P ) and 2 ρ ( r CP ) are the charge density and its Laplacian at the corresponding critical points.
Bondγ = 0.10288γ = 0.12336
L ( Å ) ρ ( r C P ) 2 ρ ( r C P ) L ( Å ) ρ ( r C P ) 2 ρ ( r C P )
B1–C1(d1)1.8210.733−4.7172.0610.461−0.606
B2–C1(d2)1.5791.128−1.8001.5401.254−1.906
C1–C2(d3)1.5001.731−16.1791.4851.791−17.796
B2–C2(d4)1.8210.733−4.7462.0610.461−0.606

Share and Cite

MDPI and ACS Style

Zheng, B.; Zhang, M.; Chang, S. Mechanical Properties and Atomic Explanation of Plastic Deformation for Diamond-Like BC2. Materials 2016, 9, 514. https://doi.org/10.3390/ma9070514

AMA Style

Zheng B, Zhang M, Chang S. Mechanical Properties and Atomic Explanation of Plastic Deformation for Diamond-Like BC2. Materials. 2016; 9(7):514. https://doi.org/10.3390/ma9070514

Chicago/Turabian Style

Zheng, Baobing, Meiguang Zhang, and Shaomei Chang. 2016. "Mechanical Properties and Atomic Explanation of Plastic Deformation for Diamond-Like BC2" Materials 9, no. 7: 514. https://doi.org/10.3390/ma9070514

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