Next Article in Journal
Fabrication and Physical Evaluation of Gelatin-Coated Carbonate Apatite Foam
Next Article in Special Issue
Mechanical and Electronic Properties of XC6 and XC12
Previous Article in Journal
Optimization of Injection Molding Parameters for HDPE/TiO2 Nanocomposites Fabrication with Multiple Performance Characteristics Using the Taguchi Method and Grey Relational Analysis
Previous Article in Special Issue
Crystal Structures and Mechanical Properties of Ca2C at High Pressure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Superhard Phase and Physical Properties of ZrB3 from First-Principles Calculations

1
College of Physics and Optoelectronics Technology, Baoji University of Arts and Sciences, Baoji 721016, China
2
College of Mathematics and Information Science, Baoji University of Arts and Sciences, Baoji 721013, China
3
School of Science, Sichuan University of Science and Engineering, Zigong 643000, China
*
Author to whom correspondence should be addressed.
Materials 2016, 9(8), 703; https://doi.org/10.3390/ma9080703
Submission received: 25 July 2016 / Revised: 11 August 2016 / Accepted: 15 August 2016 / Published: 22 August 2016
(This article belongs to the Special Issue Computational Multiscale Modeling and Simulation in Materials Science)

Abstract

:
Using the first-principles particle swarm optimization algorithm for crystal structural prediction, we have predicted a novel monoclinic C2/m structure for ZrB3, which is more energetically favorable than the previously proposed FeB3-, TcP3-, MoB3-, WB3-, and OsB3-type structures in the considered pressure range. The new phase is mechanically and dynamically stable, as confirmed by the calculations of its elastic constants and phonon dispersion curve. The calculated large shear modulus (227 GPa) and high hardness (42.2 GPa) show that ZrB3 within the monoclinic phase is a potentially superhard material. The analyses of the electronic density of states and chemical bonding reveal that the strong B–B and B–Zr covalent bonds are attributed to its high hardness. By the quasi-harmonic Debye model, the heat capacity, thermal expansion coefficient and Grüneisen parameter of ZrB3 are also systemically investigated.
PACS:
61.66.Fn; 61.50.Ah; 62.20.de

Graphical Abstract

1. Introduction

Superhard materials are considerably used in various industrial applications due to their superior properties of low compressibility, thermal conductivity, refractive index, chemical inertness, and high hardness. Previously, it was generally accepted that the superhard materials are strong covalently bonded compounds formed by light atoms B, C, N, and O, such as diamond [1], c-BN [2], B6O [3], etc. These materials are expensive and rare, for they can be synthesized only under high-temperature and high-pressure conditions. Recently, it has been reported that the incorporation of light atoms (B, C, N and O) into heavy transition metals (TMs) with high valence electron densities provides a good candidate for designing new hard materials. The compounds formed by TM and light atoms usually possess high valence electron density and directional covalent bonds, and these covalent bonds are strong enough to inhibit creation and movement of dislocations, which greatly improve their mechanical properties and create high hardness. Based on this design criterion, the recent design of new intrinsically potential superhard materials has focused on TM borides, e.g., ReB2 [4], OsB2 [5], WB4 [6,7], CrB4 [8] and so on. The obtained results exhibit that these materials all have large bulk and shear moduli. Moreover, TM borides can be synthesized under ambient pressure, which results in a low-cost synthesis condition and is beneficial to their application. Thereby, these pioneering studies open up a new route for pursuing new superhard materials.
Up to now, in the Zr-B system, three binary phases [9,10], i.e., ZrB, ZrB2, and ZrB12, have been synthesized experimentally. Among them, ZrB with the NaCl-type cubic structure has a high melting point and thermal conductivity, and its thermal stability is in the temperature range from 1073 to 1523 K [9]. ZrB2 is an AlB2-type hexagonal structure and has properties of both ceramics and metals, and its melting temperature has been reported to be either 3313 [9] or 3523 K [11,12]. Meanwhile, ZrB2 also possesses high hardness, electrical conductivity, thermal conductivity, and chemical corrosion resistance [13,14]. ZrB2 ceramic products have been considerably used in making high-temperature structural and functional materials [15,16]. ZrB12 with an UB12-type cubic structure was formed by the peritectic reaction at T = 2303 K [11], and it presents a wide range of interesting characteristics such as high temperature resistance, superconductivity, and large hardness. The crystal structure and related physical properties of ZrB, ZrB2, and ZrB12 have been investigated theoretically [17,18,19,20,21,22,23,24,25,26]. By using the first-principles calypso algorithm for crystal structure prediction, Zhang et al. [27] predicted two orthorhombic Cmcm and Amm2 structures of ZrB4, and they also showed that these two phases are potentially superhard materials because of large shear modulus (229 and 236 GPa) and high hardness (42.8 and 42.6 GPa). Very recently, Chen et al. [28] systemically investigated the elasticity, hardness, and thermal properties of ZrBn (n = 1, 2, 12) by using the first-principles calculations of plane-wave ultra-soft pseudo-potential technology based on the density functional theory. However, there is no experimental report on the synthesis of zirconium triborides; besides, the theoretical research works of zirconium triborides are also few so far.
In this work, the crystal structures of ZrB3 are extensively investigated by the first-principles particle swarm optimization algorithm (PSO) on crystal structural prediction [29,30,31]. A stable monoclinic C2/m structure has been uncovered for ZrB3, which is much more energetically preferable than the earlier proposed FeB3-, TcP3-, MoB3-, WB3-, and OsB3-type structures in the pressure range of 0–100 GPa. The lattice parameters, total energy, formation enthalpy, phonon frequency, phase stability, elastic properties, density of states, and electronic localization function are then performed to study this novel monoclinic phase by using the first-principles calculations. To further study ZrB3, the thermodynamic properties are also investigated by the quasi-harmonic Debye model.

2. Computational Methods

To search for potential structure, the PSO technique in crystal structure analysis using the CALYPSO code [32] has been implemented at 0 GPa with one to four formula units (f.u.) in each simulation cell. The underlying ab initio calculations are performed using density functional theory within the Perdew–Burke–Ernzerhof (PBE) generalized gradient approximation (GGA) of the exchange-correlation energy, as implemented in the Vienna ab initio simulation package (VASP) [33]. The all-electron projector augmented wave (PAW) method [34] is employed with 2s22p1 and 4s24p64d25s2 treated as the valence electrons for B and Zr, respectively. Geometry optimization is performed using the conjugate gradient algorithm method with a plane-wave cutoff energy of 520 eV. The calculations are conducted with 11 × 7 × 4, 9 × 11 × 6, 3 × 11 × 10, 8 × 8 × 4, 8 × 8 × 5, and 13 × 13 × 7 for the predicted C2/m-ZrB3 structure and considered FeB3-type (No, 11, P21/m), TcP3-type (No, 62, Pnma), MoB3-type (No, 166, R-3m), WB3-type (No, 194, P63/mmc), and OsB3-type (No, 187, P-6m2) structures. For hexagonal structures, Γ-centered k mesh is used, and for other structures, appropriate Monkhorst-Pack k meshes [35] are used to ensure that all the total energy calculations are well converged to better than 1 meV/atom. The phonon calculation is carried out by using a supercell approach as implemented in the PHONOPY code [36,37]. Single-crystal elastic constants are calculated by a strain-energy approach, i.e., applying a small strain to the equilibrium lattice and fitting the dependence of the resulting change in energy on the strain. The bulk modulus, shear modulus, Young’s modulus, and Poisson’s ratio are derived from the Voigt–Reuss–Hill approximation [38]. The theoretical Vickers hardness is estimated by the empirical formula proposed by Chen et al. [39]. Moreover, the quasi-harmonic Debye model [40], which is constructed from the Helmholtz free energy at the temperature below the melting point in the quasi-harmonic approximation, is used to obtain thermodynamic properties of ZrB3.

3. Results and Discussion

Through the PSO technique, we perform a variable-cell structure prediction simulation for ZrB3 including 1–4 f.u. in the simulation cell at 0 GPa, and we successfully predict a novel monoclinic structure with space group C2/m, as depicted in Figure 1. For the monoclinic C2/m structure, it contains four ZrB3 in a unit cell with the equilibrium lattice parameters of a = 3.163 Å, b = 5.440 Å, c = 8.773 Å, α = γ = 90°, and β = 93.633°, in which three inequivalent Zr, B1, and B2 atoms occupy the Wyckoff 4i (0.4677, 0, 0.2994), 8j (0.8915, 0.1694, 0.0932), and 4h (0, 0.1668, 0.5) sites, respectively. Figure 1a presents the polyhedral view for the predicted C2/m structure, which shows an intriguing B–Zr–B sandwich stacking order along the crystallographic c-axis. In this structure, B atoms form parallel hexagonal planes, and each Zr atom is coordinated by 12 neighboring B atoms, forming edge-shared ZrB12 hexagonal columns which are connected by boron hexagonal planes. For each ZrB12, the calculated B–B distances are 1.809 (×4), 1.844 (×2), 1.822 (×4), and 1.816 (×2) Å, while the calculated Zr–B distances are 2.729 (×2), 2.563 (×2), 2.434 (×2), 2.472 (×2), 2.528 (×2), and 2.599 (×2) Å.
At zero temperature, a stable crystalline structure requires all phonon frequencies to be positive. Therefore, we calculate the phonon dispersion curves for the predicted C2/m-ZrB3 at 0 and 100 GPa, respectively. As shown in Figure 2, no imaginary phonon frequency appears in the whole Brillouin zone, indicating its dynamic stability at ambient pressure and high pressure. It is well known that the thermodynamic stability of a compound can be indicated by the energy of its most stable phase. Thus, we calculate the total energy per f.u. as a function of the volume for the predicted C2/m structure, as shown in Figure 3a. For comparison, the previously known five structures of FeB3, TcP3, MoB3, WB3, and OsB3 are also considered for ZrB3. As seen from Figure 3a, our predicted C2/m structure for ZrB3 has a lower energy minimum than the FeB3-, TcP3-, MoB3-, WB3-, and OsB3-type structures. This implies that the predicted C2/m structure is the ground-state phase at 0 GPa. Figure 3b shows the enthalpy curves of ZrB3 with six different structures relative to the FeB3-ZrB3 phase. One can see clearly that the predicted C2/m phase is much more energetically favorable than the previously proposed FeB3-, TcP3-, MoB3-, WB3-, and OsB3-type structures in the whole pressure range.
We also calculate the formation enthalpies of the considered structural candidates of ZrB3 in the pressure range of 0–100 GPa. The formation enthalpy of ZrB3 with respect to the separate phases can be examined by following the reaction route Δ H = H ZrB 3 H Zr 3 H B , where ∆H represents the formation enthalpy and the hexagonal Zr and α-B are chosen as the reference phases. Figure 4a presents the calculated formation enthalpies of ZrB3 with different structures under pressure. It can be clearly seen from this figure that the predicted C2/m structure is the most stable phase of all the considered structures up to 100 GPa. All the considered structures are thermodynamically stable with a negative value of the formation enthalpy below 100 GPa. Among them, the predicted C2/m phase has the lowest formation enthalpy of −2.715 eV at 0 GPa. Therefore, the C2/m phase is more easily synthesized at ambient conditions, and meanwhile this also indicates that the predicted C2/m phase is more stable than the reference phase mentioned above. Figure 4b presents the relative enthalpy-pressure curves of C2/m-ZrB3 and its respective competing phases with respect to elemental Zr and B. Here we choose ZrB, ZrB2, and ZrB12 as the competing phases because they have been synthesized in experiments and are thermodynamically stable. From Figure 4b, the predicted C2/m-ZrB3 is much more energetically stable for decomposing into elements (Zr + B), compounds ZrB + B and ZrB12 + Zr in the pressure range from 0 to 100 GPa, whereas the competing phase ZrB2 + B is the most stable phase within the given pressure range. It is noteworthy that the relative enthalpy difference between the C2/m-ZrB3 and ZrB2 + B becomes smaller and smaller with the increasing pressure. This indicates that the ZrB3 may be the most stable phase against which to decompose into compounds ZrB2 + B when the pressure exceeds 100 GPa.
Mechanical stability is a necessary condition for the existence of a crystal, and the mechanical properties (elastic constants and elastic moduli) are important for potential technological and industrial applications. Accurate elastic constants help to understand the mechanical properties and also provide useful information for estimating the hardness of a material. By using a strain-energy method, we obtain the zero-pressure elastic constants (Cij) of the C2/m-ZrB3. The calculated elastic constants are listed in Table 1 along with the theoretical values and available experimental data of other zirconium borides, MoB3, WB3, RuB3, and OsB3. For a stable monoclinic crystal, the independent elastic stiffness tensor consists of 13 components, C11, C22, C33, C44, C55, C66, C12, C13, C23, C15, C25, C35, and C46, and its mechanical stability criteria can be given by:
C 11 > 0 , C 22 > 0 , C 33 > 0 , C 44 > 0 , C 55 > 0 , C 66 > 0 , [ 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 , 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 , { 2 [ C 15 C 25 ( C 33 C 12 C 13 C 23 ) + C 15 C 35 ( C 22 C 13 C 12 C 23 ) + 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.
As shown in Table 1, the elastic constants of the predicted C2/m-ZrB3 completely fulfill the elastic stability criteria for a monoclinic crystal, indicating that it is mechanically stable at ambient conditions. The values of C11, C22, and C33 for the new phase are all larger than 400 GPa, suggesting its strong incompressibility along the a, b, and c axis, respectively. Additionally, the values of C11 and C22 of the predicted C2/m phase are much higher than that of C33, which reflects that the bond strengths along the [100] and [010] directions are much stronger than that of the [001] direction. This is because the covalence of the B–B bond is higher than that of the Zr–B bond, as indicated by Figure 8. Thus, the relatively small C33 with regard to C11 and C12 is understandable. C44 is an important indicator for the hardness of a material. The large C44 value (207 GPa) of the predicted C2/m-ZrB3 indicates its relatively strong strength against the shear deformation. Using the obtained elastic constants, the polycrystalline bulk modulus B and shear modulus G are thus determined by the Voigt–Reuss–Hill approximation. The Young’s modulus E and Poisson’s ratio v are derived by the following formulas: E = 9BG/(3B + G) and v = (3B − 2G)/(6B + 2G). The calculated bulk modulus, shear modulus, Young’s modulus, and Poisson’s ratio of the C2/m phase together with the reference materials mentioned above are listed in Table 1. As seen from Table 1, the predicted C2/m phase has a larger bulk modulus of 238 GPa, which is larger than those of ZrB [17,18,19,28,41] and ZrB12 [26,28,43] but comparable to those of ZrB2 [17,20,21,28,42], ZrB4 [27], and other TMB3 (TM = Mo, W, Ru, Os) [45,46,47,48]. This indicates that the C2/m phase is difficult to compress. Moreover, the bulk modulus (B = 238 GPa) for C2/m-ZrB3 agrees well with that directly obtained from the fitting results (B0 = 237 GPa) of the third-order Birch-Murnaghan equation of states, which further verifies the reliability of our elastic calculations. To compare the incompressibility of the C2/m-ZrB3, other zirconium borides, c-BN, and diamond under pressure, the volume compressions as a function of pressure are shown in Figure 5. Clearly, all the considered zirconium borides except for ZrB have almost the same incompressibility due to their very close bulk moduli, but their incompressibility is lower than those of c-BN and diamond. The shear modulus of a material is a measure of the ability to resist shape change at a constant volume and it plays an important role in hardness compared with the bulk modulus. Interestingly, the shear modulus of the predicted C2/m-ZrB3 is almost the same as those of ZrB2 [17,20,21,28,42] and ZrB4 [27], suggesting that it is a potential superhard material. Except the bulk modulus and shear modulus, the Young’s modulus can also provide a good measure of the stiffness of materials. Generally speaking, the larger the Young’s modulus a material has, the harder it is to deform. Since the Young’s modulus of the C2/m-ZrB3 is similar to those of the ZrB2 [17,20,21,28,42] and ZrB4 [27], it is conceivable that the C2/m-ZrB3 is a superhard material. Further, Poisson’s ratio v is a crucial parameter to describe the degree of directionality for the covalent bonding. From Table 1, the v value of the C2/m-ZrB3 (0.14) is a little larger than that of ZrB2 [42], the same as ZrB4 [27], and much smaller than those of the other TM borides mentioned above [45,46,47,48], which means there is a strong degree of covalent bonding due to the presence of the planar six-membered ring boron network. To describe the brittleness or ductility of materials, Pugh [49] proposed an experimental criterion by the ratio G/B. The critical value which separates ductile and brittle materials is about 0.57. If G/B < 0.57, the material behaves in a ductile manner; otherwise, the material behaves in a brittle way. It can be seen from Table 1 that the G/B value for the predicted phase is 0.95, implying it is rather brittle nature. All the results strongly support that the C2/m-ZrB3 is a potential candidate for an ultra-incompressible and superhard material.
The Debye temperature is a very important parameter of materials, and it is closely related to many physical properties such as specific heat, elastic constants, melting temperature, hardness and so on [50]. It is used to differentiate between high- and low-temperature regions for a solid. If the temperature T > ΘD, all modes are expected to have the energy of kBT; if T < ΘD, the high-frequency modes are expected to be frozen, namely the vibrational excitations originate only from the acoustic vibrations. The Debye temperature ΘD for our studied C2/m-ZrB3 is estimated from the average sound velocity (vm) by the following expression:
Θ D = h k [ 3 n 4 π ( ρ N A M ) ] 1 3 v m
where h is Planck’s constant, k is Boltzmann’s constant, NA is Avogadro’s number, n is the number of atoms per formula unit, M is the molecular mass per f.u., and ρ is the density. The average sound velocity vm is given by:
v m = [ 1 3 ( 2 v t 3 + 1 v l 3 ) ] 1 3
where vt and vl are the transverse and longitudinal elastic wave velocities of the polycrystalline material, which is determined by Anderson’s method [51]. Figure 6 presents the dependence of the Debye temperatures of the C2/m-ZrB3 on the pressure. At zero pressure and zero temperature, our calculated Debye temperature is 998 K for the C2/m-ZrB3, which is much larger than the values of the known ultra-incompressible RuB2 (780 K) [52], ReN2 (735 K) [53], ReO2 (850 K) [54], and ReB2 (858.3 K) [55]. Generally, the higher the Debye temperature materials possess, the larger their microhardness. Clearly, the Debye temperature becomes greater with increasing the pressure. This indicates that the pressure is in favor of the improvement of the hardness for the C2/m-ZrB3.
Due to the high bulk and shear moduli as well as the large Debye temperature for the predicted C2/m-ZrB3, the hardness calculation is of great importance. The hardness of a material is the intrinsic resistance to deformation when a force is loaded, which depends upon the loading force and the sample quality. The Vickers hardness of a material is estimated by [39]:
H v = 2 ( K 2 G ) 0.585 3
where G is the shear modulus and K = G/B is the Pugh modulus ratio. Using this model, the estimated Vickers hardness of the C2/m-ZrB3 is 42.2 GPa, which exceeds the lower limit of superhard materials (40 GPa) and is comparable to the known superhard materials B6O (45 GPa) [3] and c-BN (48 GPa) [56]. Therefore, the C2/m-ZrB3 is considered to be a superhard material.
The electronic structure and chemical bonding are key factors to further understand the origin of the mechanical properties of the C2/m-ZrB3. For this purpose, the density of states (DOS) and bond characteristics are calculated, and the corresponding results are analyzed here. Figure 7a presents the total and partial density of states of the C2/m-ZrB3 at 0 GPa, where the vertical dashed line denotes the Fermi level (EF). As seen from this figure, the C2/m-ZrB3 shows a metallic behavior because of the finite electronic DOS at the Fermi level. From the partial DOS, it exhibits that the peaks below −11 eV are mainly attributed to B-s and B-p states with slight contributions from Zr-s, Zr-p, and Zr-d states. The states above −11 eV largely come from Zr-d and B-p orbitals with small contributions of B-s, Zr-s, and Zr-p. Moreover, the partial DOS profiles of Zr-d and B-p have a very similar shape in the range from −11 to 0 eV, which indicates the significant hybridization between these two orbitals. This hybridization also reflects a strong covalent interaction between the Zr and B atoms. On the other hand, the DOS profile near EF mainly originates from the 4d state of Zr. Another typical feature of DOS is that there is a deep valley, namely the pseudogap around EF, which is regarded as the borderline between the bonding and antibonding states [57,58,59]. Significantly, the pseudogap appears below the EF in the C2/m-ZrB3, indicating the s-p and p-d bonding states started to be saturated. The nearly full occupation of the bonding states, and without filling the antibonding states, results in the high bulk modulus, large shear modulus, and small Poisson’s ratio, and also increases the structural stability of the C2/m-ZrB3. To further understand the changes of the total and partial DOS under pressure, Figure 7b–d present the total and partial DOS of the C2/m-ZrB3 at 20, 50, and 100 GPa, respectively. Compared with the case for 0 GPa (see Figure 7a), the shapes of the total and partial DOS are slightly changed with the increase of the pressure. This indicates that the C2/m-ZrB3 structure is still stable under pressure up to 100 GPa. Traditionally, the stability of a solid is closely associated with the DOS at the Fermi energy. According to our calculations, the DOS values at the Fermi level are 0.294 eV for 0 GPa, 0.263 eV for 20 GPa, 0.227 eV for 50 GPa, and 0.186 eV for 100 GPa. Therefore, the DOS value at the Fermi level decreases with increasing the pressure. Based on the previous reference report [60], there is a principle stating that the DOS at the Fermi level would be hopefully smaller in an energetically-favored structure. This means that the applied pressure is in favor of the stability of ZrB3. In addition, one can see from Figure 7 that the total DOS reveals a slight broadening with the enhancement of the pressure. All these results may be explained by the variations of the spacing between atoms.
To gain deeper insight into the bonding character of the predicted C2/m-ZrB3, we calculate the electronic localization function (ELF), which is based upon a topological analysis related to the Pauli exclusion principle. The ELF is a contour plot in real space, in which different contours have values from 0 to 1. The upper limit ELF = 1 shows the perfect localization property of covalent bonds or lone pairs (filled core levels), whereas the lower limit ELF = 0 is typical for a vacuum (no electron density) or areas between atomic orbitals. ELF = 0.5 corresponds to the perfect free-electron gas, with values of this order meaning regions with bonding of a metallic character. Note that ELF is not a measure of electron density but a measure of the Pauli principle, and it is used to distinguish metallic, covalent, and ionic bonding. Figure 8 presents the calculated ELF contours of the predicted C2/m-ZrB3 phase on the (001) and (102) planes. Because of the high ELF value between the two adjacent B and B atoms, as shown in Figure 8a, there is the existence of a strong covalent B–B bonding within the planar six-membered ring unit. Meanwhile, one can see from Figure 8b that the large ELF value between the Zr and B atoms implies the partially B–Zr covalent bonding interaction in the C2/m-ZrB3 phase. Therefore, the strong covalent interaction between B–B bonds and B–Zr bonds is the major reason for its high hardness and stability.
The investigation on the thermodynamic properties of solids at high temperature and high pressure is an interesting topic in materials science. Therefore, we investigate the thermodynamic properties of the C2/m-ZrB3 in the temperature range from 0 to 1500 K, where the quasi-harmonic model remains fully valid and the pressure effect is also investigated in the pressure range from 0 to 100 GPa. By applying this model, we can get the variations of the lattice heat capacity CV or CP, the thermal expansion coefficient α, and Grüneisen parameter γ with temperature and pressure. Figure 9 presents the temperature dependences of the heat capacity at constant volume CV and the heat capacity at constant pressure CP at different pressures, respectively. As seen from Figure 9, when the temperature is lower than 400 K, the difference of between CV and CP is very small. This can be explained by the relation between CP and CV, i.e., CP = CV + TVBα2 [61], since the difference between the CP and CV is mainly determined by α2 for the case of low temperature. Therefore, when the thermal expansion coefficient α becomes small, e.g., the temperature T decreases and the pressure P increases, the CP-CV difference also becomes small. On the other hand, the CV and CP increase rapidly with the temperature at low temperature. This rapid increase is simply due to the exponentially increased number of excited phonon modes. However, when the applied temperature is considerably high, due to the anharmonic effect, the CP is different from the CV. The former is proportional to T at high temperature, while the latter reaches a constant value which is the so-called Dulong-Petit limit (CV(T) ~ 3R for monoatomic solids). Moreover, as the pressure keeps constant, the CV and CP increase with the temperature. As the temperature keeps constant, CV and CP decrease with the pressure. This shows that the temperature has a more significant influence on the heat capacity than the pressure.
Figure 10 plots the changes of the thermal expansion coefficient α for the C2/m-ZrB3 with temperature and pressure. As shown in Figure 10a, the thermal expansion coefficient α increases sharply at low temperatures, especially for the case of 0 GPa, and then gradually approaches a linear increase at high temperature, finally becomes moderate at much higher temperature. In terms of α ~ CV/B [61], the fast increase of α at low temperature is mainly due to that of CV, and the linear increase of α at high temperature is attributed to that of bulk modulus B (CV is saturated to the Dulong-Petit limit at high temperature). In addition, the effect of the pressure on α is very small for the case of low temperature, whereas the effect is clearly enhanced at sufficiently high temperature. When the pressure increases, α reduces drastically for a given temperature and the effect of the temperature on it becomes unnoticeable. From Figure 10b, one can see clearly that α is almost close to a constant value at high temperature and pressure. All these results are in good agreement with those of many kinds of materials by applying the Debye theory, such as OsB4, FeB4, and TcN [62,63,64].
The Grüneisen parameter γ can describe the anharmonic effects in the vibration of a crystal lattice, and it is commonly used to characterize and explore the thermodynamic behavior of a material at high temperature and high pressure, such as the thermal expansion coefficient and the temperature dependence of phonon frequencies and line-widths. For this purpose, we investigate the dependences of the Grüneisen parameter γ for the C2/m-ZrB3 with the temperature and pressure, and the corresponding results are given in Figure 11. As seen from Figure 11a, for a given pressure, γ almost remains the same at the low temperature of ≤300 K. As the applied temperature is higher than 300 K, γ tonelessly increases with temperature for the given pressure. In addition, one can find from Figure 11b that γ decreases nearly exponentially with the pressure for a given temperature. As the applied pressure increases, the temperature has a much weaker influence on γ than the pressure.

4. Conclusions

In summary, a novel monoclinic C2/m structure is unraveled to be the ground-state structure for ZrB3 via the PSO algorithm combined with first-principles calculations. The phonon dispersion and elastic constant calculations have verified that the C2/m-ZrB3 is dynamically and mechanically stable. The formation enthalpy-pressure relationship shows that this new phase is energetically superior to the previously proposed FeB3-, TcP3-, MoB3-, WB3-, and OsB3-type structures within the pressure range of 0–100 GPa. In addition, the high bulk modulus, large shear modulus, and small Poisson’s ratio of the predicted C2/m phase indicate that it is a promising low-compressibility material. Based on the calculated Vickers hardness (42.2 GPa) for the C2/m phase, it is a potential superhard material. The more detailed analyses of the electronic structure and electronic localization function have further demonstrated that the strong covalent B–B bonding and B–Zr bonding are the main driving force for the incompressibility and hardness for C2/m-ZrB3. By using the quasi-harmonic Debye model, some fundamental thermodynamic properties, such as the heat capacity, thermal expansion coefficient, and Grüneisen parameter are estimated in the pressure range of 0–100 GPa and the temperature range of 0–1500 K. We hope that the present theoretical work will stimulate further experimental research on this material in the future.

Acknowledgments

This work was supported by the Science Foundation of Baoji University of Arts and Sciences of China (Grant Nos. ZK16069 and ZK11135), the Natural Science Foundation of the Education Committee of Shaanxi Province of China (Grant No. 2013JK0637), the Natural Science Basic Research plan in Shaanxi Province of China (Grant No. 2016JM1012), and the Natural Science Foundation of China (Grant No. 11404008).

Author Contributions

Gangtai Zhang and Yanfei Hu designed the project; Gangtai Zhang and Tingting Bai performed the calculations; Gangtai Zhang and Tingting Bai wrote the manuscript; Yaru Zhao and Yanfei Hu revised the paper and coordinated the work; all authors discussed results.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Occelli, F.; Farber, D.L.; Toullec, R.L. Properties of diamond under hydrostatic pressures up to 140 GPa. Nat. Mater. 2003, 2, 151–154. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, Y.; Sun, H.; Chen, C.F. Structural deformation, strength and instability of cubic BN compared to diamond: A first-principles study. Phys. Rev. B 2006, 73, 144115. [Google Scholar] [CrossRef]
  3. He, D.W.; Zhao, Y.S.; Daemen, L.; Qian, J.; Shen, T.D.; Zerda, T.W. Boron suboxide: As hard as cubic boron nitride. Appl. Phys. Lett. 2002, 81, 643–645. [Google Scholar] [CrossRef]
  4. Chung, H.Y.; Weinberger, M.B.; Levine, J.B.; Kavner, A.; Yang, J.M.; Tolbert, S.H.; Kaner, R.B. Synthesis of ultra-incompressible superhard rhenium diboride at ambient pressure. Science 2007, 316, 436–439. [Google Scholar] [CrossRef] [PubMed]
  5. Cumberland, R.W.; Weinberger, M.B.; Gilman, J.J.; Clark, S.M.; Tolbert, S.H.; Kaner, R.B. Osmium diboride, an ultra-incompressible, hard material. J. Am. Chem. Soc. 2005, 127, 7264–7265. [Google Scholar] [CrossRef] [PubMed]
  6. Gu, Q.F.; Krauss, G.; Steurer, W. Transition metal borides: Superhard versus ultra-incompressible. Adv. Mater. 2008, 20, 3620–3626. [Google Scholar] [CrossRef]
  7. Zhang, R.F.; Legut, D.; Lin, Z.J.; Zhao, Y.S.; Mao, H.K.; Veprek, S. Stability and strength of transition-metal tetraborides and triborides. Phys. Rev. Lett. 2012, 108, 255502. [Google Scholar] [CrossRef] [PubMed]
  8. Gou, H.Y.; Li, Z.P.; Niu, H.; Gao, F.M.; Zhang, J.W.; Rodney, C.E.; Lian, J. Unusual rigidity and ideal strength of CrB4 and MnB4. Appl. Phys. Lett. 2012, 100, 111907. [Google Scholar] [CrossRef]
  9. Glaser, F.W.; Post, B. Phase diagram zirconium–boron. Trans. AIME 1953, 197, 1117–1118. [Google Scholar]
  10. Nowotny, H.; Rudy, E.; Benesovsky, F. Untersuchungen in den systemen: Zirkonium-bor-kohlenstoff und zirkonium-bor-stickstoff. Mh. Chem. 1960, 91, 963–974. (In Germany) [Google Scholar] [CrossRef]
  11. Portnoi, K.I.; Romashov, V.M.; Burobina, L.N. Constitution diagram of the system zirconium-boron. Sov. Powder Metall. Met. Ceram. 1970, 9, 577–580. [Google Scholar] [CrossRef]
  12. Tokunaga, T.; Terashima, K.; Ohtani, H.; Hasebe, M. Thermodynamic analysis of the phase equilibria in the Fe-Zr-B system. Mater. Trans. 2008, 49, 2534–2540. [Google Scholar] [CrossRef]
  13. Opeka, M.M.; Talmy, I.G.; Wuchina, E.J.; Zaykoski, J.A.; Causey, S.J. Mechanical, thermal, and oxidation properties of refractory hafnium and zirconium compounds. J. Eur. Ceram. Soc. 1999, 19, 2405–2414. [Google Scholar] [CrossRef]
  14. Wang, Y.G.; Liu, W.; Cheng, L.F.; Zhang, L.T. Preparation and properties of 2D C/ZrB2-SiC ultra high temperature ceramic. Mater. Sci. Eng. A 2009, 524, 129–133. [Google Scholar] [CrossRef]
  15. Cheng, Y.M.; Gadalla, A.M. Synthesis and analysis of ZrB2-based composites. Mater. Manuf. Process. 1996, 11, 575–587. [Google Scholar] [CrossRef]
  16. Martinez, M.; Rodriguez, D. Characterisation and high temperature mechanical properties of zirconium boride-based materials. J. Eur. Ceram. Soc. 2002, 22, 2543–2549. [Google Scholar] [CrossRef]
  17. Li, H.; Zhang, L.T.; Zeng, Q.F.; Wang, J.J.; Cheng, L.F.; Ren, H.T.; Guan, K. Crystal structure and elastic properties of ZrB compared with ZrB2: A first-principles study. Comput. Mater. Sci. 2010, 49, 814–819. [Google Scholar] [CrossRef]
  18. Wang, Y.; Chen, W.; Chen, X.; Liu, H.Y.; Ding, Z.H.; Ma, Y.M.; Wang, X.D.; Cao, Q.P.; Jiang, J.Z. Crystal structures, stability, electronic and elastic properties of 4d and 5d transition metal monoborides: First-principles calculations. J. Alloys Compd. 2012, 538, 115–124. [Google Scholar] [CrossRef]
  19. Huang, B.; Duan, Y.H.; Hu, W.C.; Sun, Y.; Chen, S. Structural, anisotropic elastic and thermal properties of MB (M = Ti, Zr and Hf) monoborides. Ceram. Int. 2015, 41, 6831–6843. [Google Scholar] [CrossRef]
  20. Okamoto, N.L.; Kusakari, M.; Tanaka, K.; Inui, H.; Yamaguchi, M.; Otani, S. Temperature dependence of thermal expansion and elastic constants of single crystals of ZrB2 and the suitability of ZrB2 as a substrate for GaN film. J. Appl. Phys. 2003, 93, 88–93. [Google Scholar] [CrossRef] [Green Version]
  21. Zhang, X.H.; Luo, X.G.; Han, J.C.; Li, J.P.; Han, W.B. Electronic structure, elasticity and hardness of diborides of zirconium and hafnium: First principles calculations. Comput. Mater. Sci. 2008, 44, 411–421. [Google Scholar] [CrossRef]
  22. Zhang, X.H.; Luo, X.G.; Li, J.P.; Han, J.C.; Han, W.B.; Hong, C.Q. Structure and bonding features of ZrB2 (0001) surface. Comput. Mater. Sci. 2009, 46, 1–6. [Google Scholar] [CrossRef]
  23. Okamoto, N.L.; Kusakari, M.; Tanaka, K.; Inui, H.; Otani, S. Anisotropic elastic constants and thermal expansivities in monocrystal CrB2, TiB2, and ZrB2. Acta Mater. 2010, 58, 76–84. [Google Scholar] [CrossRef] [Green Version]
  24. Jäger, B.; Paluch, S.; Żogal, O.J.; Wolf, W.; Herzig, P.; Filippov, V.B.; Shitsevalova, N.; Paderno, Y. Characterization of the electronic properties of YB12, ZrB12, and LuB12 using 11B NMR and first-principles calculations. J. Phys. Condens. Matter 2006, 18, 2525–2535. [Google Scholar] [CrossRef]
  25. Korozlu, N.; Colakoglu, K.; Deligoz, E.; Aydin, S. The elastic and mechanical properties of MB12 (M = Zr, Hf, Y, Lu) as a function of pressure. J. Alloys Compd. 2013, 546, 157–164. [Google Scholar] [CrossRef]
  26. Ai, B.C.; Luo, X.G.; Yu, J.J.; Miao, W.B.; Hu, P. Theoretical elastic stiffness and thermodynamic properties of zirconium dodecaboride from first principles calculation. Comput. Mater. Sci. 2014, 82, 37–44. [Google Scholar] [CrossRef]
  27. Zhang, X.Y.; Qin, J.Q.; Sun, X.W.; Xue, Y.N.; Ma, M.Z.; Liu, R.P. First-principles structural design of superhard material of ZrB4. Phys. Chem. Chem. Phys. 2013, 15, 20894–20899. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, Z.Q.; Peng, Y.S.; Hu, M.; Li, C.M.; Luo, Y.T. Elasticity, hardness, and thermal properties of ZrBn (n = 1, 2, 12). Ceram. Int. 2016, 42, 6624–6631. [Google Scholar] [CrossRef]
  29. Wang, Y.C.; Lv, J.; Zhu, L.; Ma, Y.M. Crystal structure prediction via particle-swarm optimization. Phys. Rev. B 2010, 82, 094116. [Google Scholar] [CrossRef]
  30. Wang, Y.C.; Lv, J.; Zhu, L.; Ma, Y.M. CALYPSO: A method for crystal structure prediction. Comput. Phys. Commun. 2012, 183, 2063–2070. [Google Scholar] [CrossRef]
  31. Wang, H.; Wang, Y.C.; Lv, J.; Li, Q.; Zhang, L.J.; Ma, Y.M. CALYPSO structure prediction method and its wide application. Comput. Mater. Sci. 2016, 112, 406–415. [Google Scholar] [CrossRef]
  32. Ma, Y.M.; Wang, Y.C.; Lv, J.; Zhu, L. CALYPSO (Crystal Structure AnaLYsis by Particle Swarm Optimization). Available online: http://www.calypso.cn (accessed on 18 August 2016).
  33. 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]
  34. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  35. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  36. Togo, A.; Oba, F.; Tanaka, I. First-principles calculations of the ferroelastic transition between rutile-type and CaCl2-type SiO2 at high pressures. Phys. Rev. B 2008, 78, 134106. [Google Scholar] [CrossRef]
  37. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef] [Green Version]
  38. Hill, R. The elastic behaviour of a crystalline aggregate. Proc. Phys. Soc. 1952, 65, 349–354. [Google Scholar] [CrossRef]
  39. Chen, X.Q.; Niu, H.; Li, D.; Li, Y. Modeling hardness of polycrystalline materials and bulk metallic glasses. Intermetallics 2011, 19, 1275–1281. [Google Scholar] [CrossRef]
  40. Blanco, M.A.; Francisco, E.; Luaña, V. GIBBS: Isothermal-isobaric thermodynamics of solids from energy curves using a quasi-harmonic Debye model. Comput. Phys. Commun. 2004, 158, 57–72. [Google Scholar] [CrossRef]
  41. Fahrenholtz, G.W.; Hilmas, G.E. Refractory diborides of zirconium and hafnium. J. Am. Ceram. Soc. 2007, 90, 1347–1364. [Google Scholar] [CrossRef]
  42. Wiley, D.E.; Manning, W.R.; Hunter, O. Elastic properties of polycrystalline TiB2, ZrB2 and HfB2 from room temperature to 1300 K. J. Less Common Met. 1969, 18, 149–157. [Google Scholar] [CrossRef]
  43. Grechnev, G.E.; Baranovskiy, A.E.; Fil, V.D.; Ignatova, T.V.; Kolobov, I.G.; Logosha, A.V.; Shitsevalova, N.Y.; Filippov, V.B.; Eriksson, O. Electronic structure and bulk properties of MB6 and MB12 borides. Low. Temp. Phys. 2008, 34, 921. [Google Scholar] [CrossRef]
  44. Kim, K.H.; Shim, K.B. The effect of lanthanum on the fabrication of ZrB2–ZrC composites by spark plasma sintering. Mater. Charact. 2003, 50, 31–37. [Google Scholar] [CrossRef]
  45. Wang, Y.C.; Yao, T.K.; Wang, L.M.; Yao, J.L.; Li, H.; Zhang, J.W.; Gou, H.Y. Structural and relative stabilities, electronic properties and possible reactive routing of osmium and ruthenium borides from first-principles calculations. Dalton Trans. 2013, 42, 7041–7050. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, M.G.; Wang, H.; Wang, H.B.; Cui, T.; Ma, Y.M. Structural modifications and mechanical properties of molybdenum borides from first principles. J. Phys. Chem. C 2010, 114, 6722–6725. [Google Scholar] [CrossRef]
  47. Liang, Y.C.; Yuan, X.; Zhang, W.Q. Thermodynamic identification of tungsten borides. Phys. Rev. B 2011, 83, 220102. [Google Scholar] [CrossRef]
  48. Ji, Z.W.; Hu, C.H.; Wang, D.H.; Zhong, Y.; Yang, J.; Zhang, W.Q.; Zhou, H.Y. Mechanical properties and chemical bonding of the Os–B system: A first-principles study. Acta Mater. 2012, 60, 4208–4217. [Google Scholar] [CrossRef]
  49. Pugh, S.F. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Philos. Mag. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  50. Ravindran, P.; Fast, L.; Korzhavyi, P.A.; Johansson, B.; Wills, J.; Eriksson, O. Density functional theory for calculation of elastic properties of orthorhombic crystals: Application to TiSi2. J. Appl. Phys. 1998, 84, 4891–4904. [Google Scholar] [CrossRef]
  51. Anderson, O.L. A simplified method for calculating the Debye temperature from elastic constants. J. Phys. Chem. Solids 1963, 24, 909–917. [Google Scholar] [CrossRef]
  52. Hao, X.F.; Xu, Y.H.; Wu, Z.J.; Zhou, D.F.; Liu, X.J.; Meng, J. Elastic anisotropy of OsB2 and RuB2 from first-principles study. J. Alloys Compd. 2008, 453, 413–417. [Google Scholar] [CrossRef]
  53. Li, Y.L.; Zeng, Z. New potential super-incompressible phase of ReN2. Chem. Phys. Lett. 2009, 474, 93–96. [Google Scholar] [CrossRef]
  54. Li, Y.L.; Zeng, Z. Structural, elastic and electronic properties of ReO2. Chin. Phys. Lett. 2008, 25, 4086–4089. [Google Scholar] [CrossRef]
  55. Hao, X.F.; Xu, Y.H.; Wu, Z.J.; Zhou, D.F.; Liu, X.J.; Cao, X.Q.; Meng, J. Low-compressibility and hard materials ReB2 and WB2: Prediction from first-principles study. Phys. Rev. B 2006, 74, 224112. [Google Scholar] [CrossRef]
  56. Sung, C.M.; Sung, M. Carbon nitride and other speculative superhard materials. Mater. Chem. Phys. 1996, 43, 1–18. [Google Scholar] [CrossRef]
  57. Xu, J.H.; Freeman, A.J. Phase stability and electronic structure of ScA13 and ZrA13 and of Sc-stabilized cubic ZrA13 precipitates. Phys. Rev. B 1990, 41, 12553–12561. [Google Scholar] [CrossRef]
  58. Grechnev, G.E.; Ushakova, N.V.; Kervalishvili, P.D.; Kvachantiradze, G.G.; Kharebov, K.S. Electron structure of diborides of 3d metals. Low Temp. Phys. 1997, 23, 217–219. [Google Scholar] [CrossRef]
  59. Vajeeston, P.; Ravindran, P.; Ravi, C.; Asokamani, R. Electronic structure, bonding, and ground-state properties of AlB2-type transition-metal diborides. Phys. Rev. B 2001, 63, 045115. [Google Scholar] [CrossRef]
  60. Imai, Y.; Mukaida, M.; Tsunoda, T. Comparison of density of states of transition metal disilicides and their related compounds systematically calculated by a first-principle pseudopotential method using plane-wave basis. Intermetallics 2000, 8, 381–390. [Google Scholar] [CrossRef]
  61. Huang, L.F.; Lu, X.Z.; Tennessen, E.; Rondinelli, J.M. An efficient ab-initio quasiharmonic approach for the thermodynamics of solids. Comput. Mater. Sci. 2016, 120, 84–93. [Google Scholar] [CrossRef]
  62. Yan, H.Y.; Zhang, M.G.; Huang, D.H.; Wei, Q. First-principles study of elastic and thermodynamic properties of orthorhombic OsB4 under high pressure. Solid State Sci. 2013, 18, 17–23. [Google Scholar] [CrossRef]
  63. Zhang, X.Y.; Qin, J.Q.; Ning, J.L.; Sun, X.W.; Li, X.T.; Ma, M.Z.; Liu, R.P. First principle study of elastic and thermodynamic properties of FeB4 under high pressure. J. Appl. Phys. 2013, 114, 183517. [Google Scholar] [CrossRef]
  64. Zhao, Y.R.; Zhang, G.T.; Yan, H.Y.; Bai, T.T.; Zheng, B.B.; Yuan, Y.Q. First-principles investigations of the structure and physical properties for new TcN crystal structure. Mol. Phys. 2016, 114, 1–8. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of the C2/m-ZrB3 phases. Large and small spheres represent Zr and B atoms, respectively. (a) Polyhedral view of the C2/m phase; (b) ZrB12.
Figure 1. Crystal structure of the C2/m-ZrB3 phases. Large and small spheres represent Zr and B atoms, respectively. (a) Polyhedral view of the C2/m phase; (b) ZrB12.
Materials 09 00703 g001
Figure 2. Phonon dispersion curves of the C2/m-ZrB3 at 0 GPa (a) and 100 GPa (b).
Figure 2. Phonon dispersion curves of the C2/m-ZrB3 at 0 GPa (a) and 100 GPa (b).
Materials 09 00703 g002
Figure 3. (a) Total energy versus f.u. volume for ZrB3 with six different structures; (b) Enthalpies of ZrB3 with six different structures relative to FeB3-ZrB3 phase as a function of pressure.
Figure 3. (a) Total energy versus f.u. volume for ZrB3 with six different structures; (b) Enthalpies of ZrB3 with six different structures relative to FeB3-ZrB3 phase as a function of pressure.
Materials 09 00703 g003
Figure 4. (a) The formation enthalpy-pressure curves for ZrB3 with six different structures; (b) The relative enthalpy-pressure curve of the C2/m-ZrB3 and its respective competing phases with respect to elemental Zr and B.
Figure 4. (a) The formation enthalpy-pressure curves for ZrB3 with six different structures; (b) The relative enthalpy-pressure curve of the C2/m-ZrB3 and its respective competing phases with respect to elemental Zr and B.
Materials 09 00703 g004
Figure 5. The calculated volume compression as a function of pressure for the C2/m-ZrB3 in contrast to ZrB, ZrB2, ZrB4, c-BN, and diamond.
Figure 5. The calculated volume compression as a function of pressure for the C2/m-ZrB3 in contrast to ZrB, ZrB2, ZrB4, c-BN, and diamond.
Materials 09 00703 g005
Figure 6. Debye temperature of the C2/m-ZrB3 as a function of pressure.
Figure 6. Debye temperature of the C2/m-ZrB3 as a function of pressure.
Materials 09 00703 g006
Figure 7. Total and partial density of states of the C2/m-ZrB3: (a) 0 GPa; (b) 20 GPa; (c) 50 GPa; and (d) 100 GPa. The vertical dashed line denotes the Fermi level EF.
Figure 7. Total and partial density of states of the C2/m-ZrB3: (a) 0 GPa; (b) 20 GPa; (c) 50 GPa; and (d) 100 GPa. The vertical dashed line denotes the Fermi level EF.
Materials 09 00703 g007
Figure 8. Contours of ELF of the C2/m-ZrB3 on the (001) (a) and (102) (b) planes at 0 GPa.
Figure 8. Contours of ELF of the C2/m-ZrB3 on the (001) (a) and (102) (b) planes at 0 GPa.
Materials 09 00703 g008
Figure 9. The temperature dependences of the heat capacity at constant volume CV and heat capacity at constant pressure CP at different pressures for the C2/m-ZrB3: (a) CV contours; (b) CV-T; (c) CP contours and (d) CP-T, respectively.
Figure 9. The temperature dependences of the heat capacity at constant volume CV and heat capacity at constant pressure CP at different pressures for the C2/m-ZrB3: (a) CV contours; (b) CV-T; (c) CP contours and (d) CP-T, respectively.
Materials 09 00703 g009
Figure 10. Temperature (a) and pressure (b) dependence of the thermal expansion coefficient α for the C2/m-ZrB3.
Figure 10. Temperature (a) and pressure (b) dependence of the thermal expansion coefficient α for the C2/m-ZrB3.
Materials 09 00703 g010
Figure 11. Temperature (a) and pressure (b) dependence of the Grüneisen parameter γ for the C2/m-ZrB3.
Figure 11. Temperature (a) and pressure (b) dependence of the Grüneisen parameter γ for the C2/m-ZrB3.
Materials 09 00703 g011
Table 1. The calculated elastic constants Cij (GPa), bulk modulus B (GPa), EOS fitted bulk modulus B0 (GPa), shear modulus G (GPa), Young’s modulus E (GPa), G/B ratio, Poisson’s ratio v, Debye temperature ΘD (K), and hardness Hv (GPa) for the C2/m-ZrB3.
Table 1. The calculated elastic constants Cij (GPa), bulk modulus B (GPa), EOS fitted bulk modulus B0 (GPa), shear modulus G (GPa), Young’s modulus E (GPa), G/B ratio, Poisson’s ratio v, Debye temperature ΘD (K), and hardness Hv (GPa) for the C2/m-ZrB3.
StructureWorkC11C22C33C44C55C66C12C13 C23BB0GEG/BνΘDHv
C2/m-ZrB3this work576564466207244250451151092382372275170.950.1499842.2
Fm-3m-ZrBthis work363--60--60--161160882230.550.2752910.5
theory [17]282--111--100--161-103254-0.24--
theory [18]350--43--70--163-71186-0.31-5.5
theory [19]353--51--64--160-78202-0.29-11.8
theory [28]274--42--51--125-631620.500.284507.9
experiment [41]---------148--244-0.23--
P6/mmm-ZrB2this work565-432252-25554121-2392382335270.980.1393044.1
theory [17]504-427241--91112-229-211484-0.15--
theory [21]551-436252--65121-239-229520-0.14921-
theory [28]502-374225-2294385-199-2114681.060.1188443.5
experiment [20]581-445240-26155121-245-243554----
experiment [42]---------220-225502-0.11910
Amm2-ZrB4theory [27]55457645422324325450122113241-2295220.950.14 a-42.6
Fm-3m-ZrB12this work412--251--144--2332331954580.840.1793332.5
theory [26]413--244--141--232-1934530.830.1741206-
theory [28]420--249--123--222-2024660.910.15123135.4
experiment [43]443--265--129--234----0.231260-
experiment [44]--- ----- -----1260-
Pnma-RuB3theory [45]497462471224219187111229-267 1854510.690.219-26.3
R-3m-MoB3theory [46]602-420247--106166-2762772225260.800.18-31.8
P63/mmc-WB3theory [47]656-479277-----291-252588-0.168-42
P-6m2-OsB3theory [48]525-751186--125221-317-1954850.620.245 36.9
P21/m-OsB3theory [45]67452558413529112897247-303-1864630.610.245 25.1
a Our calculated result from Reference [27] on the basis of v = (3B − 2G)/(6B + 2G).

Share and Cite

MDPI and ACS Style

Zhang, G.; Bai, T.; Zhao, Y.; Hu, Y. A New Superhard Phase and Physical Properties of ZrB3 from First-Principles Calculations. Materials 2016, 9, 703. https://doi.org/10.3390/ma9080703

AMA Style

Zhang G, Bai T, Zhao Y, Hu Y. A New Superhard Phase and Physical Properties of ZrB3 from First-Principles Calculations. Materials. 2016; 9(8):703. https://doi.org/10.3390/ma9080703

Chicago/Turabian Style

Zhang, Gangtai, Tingting Bai, Yaru Zhao, and Yanfei Hu. 2016. "A New Superhard Phase and Physical Properties of ZrB3 from First-Principles Calculations" Materials 9, no. 8: 703. https://doi.org/10.3390/ma9080703

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