Anisotropy in elasticity and thermodynamic properties of zirconium tetraboride under high pressure

Ruru Haoa, Xinyu Zhang*a, Jiaqian Qin*b, Jinliang Ninga, Suhong Zhanga, Zhi Niuc, Mingzhen Maa and Riping Liua
aState Key Laboratory of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao 066004, P. R. China. E-mail: xyzhang@ysu.edu.cn
bMetallurgy and Materials Science Research Institute, Chulalongkorn University, Bangkok 10330, Thailand. E-mail: jiaqianqin@gmail.com
cCollege of Mechanical Engineering, Yanshan University, Qinhuangdao 066004, P. R. China

Received 9th August 2015 , Accepted 4th September 2015

First published on 7th September 2015


Abstract

The recently predicted ZrB4 with an Amm2 orthorhombic structure has great scientific and technical significance owing to its novel B–Zr–B “sandwich” layer bonding and evaluated high hardness. To better understand the performance of Amm2-ZrB4, its elastic and thermodynamic properties under pressure and temperature are studied here by taking advantage of first principles calculations in combination with the quasi-harmonic Debye model. It is found that ZrB4 keeps brittleness and mechanical stability up to 100 GPa, possessing pronounced elastic anisotropy demonstrated by the elastic anisotropy factors, the direction-dependent Young's modulus, shear modulus and Poisson's ratio. The pressure and temperature dependences of the thermodynamics parameters including normalized volume V/V0, bulk modulus, specific heat, Debye temperature, thermal expansion coefficient and Grüneisen parameter in wide temperature (0–1000 K) and pressure (0–50 GPa) ranges are obtained and discussed in detail.


I. Introduction

In recent decades, transition metal borides (TMBs) have drawn considerable attention as candidate (super) hard materials, and a number of them have been widely used in high-temperature environments, cutting tools, and hard coatings owing to their superior properties such as high strength, high hardness, ultra-incompressibility and good thermal stability.1–4 Recently, novel transition metal (e.g., Fe,5–7 W,8–10 and Cr,3) borides have been successfully synthesized under ambient or high pressure (65 GPa for Pnnm-FeB4, 46.2 GPa for P63/mmc-WB4 and 48 GPa for Pnnm-CrB4) and extensive experimental and theoretical investigations have been carried out on these borides, which identifies their superhardness and leads to the low-cost synthesis of superhard materials. For the Zr–B system, there are three identified phases (ZrB, ZrB2, and ZrB12) according to the phase diagram.11 The relatively high hardness of ZrB2 and ZrB12 naturally leads us to wonder if there are any (super)hard zirconium tetraborides. Inspired by such a hypothesis, our group successfully predicted two new orthorhombic phases of ZrB4 and estimated their hardness as 42.8 GPa and 42.6 GPa for Cmcm and Amm2 structure, respectively.12 Both phases exhibit an interesting B–Zr–B sandwiches stacking order along the c and a-axis, and the sandwiches are connected by strong covalent bond (B–B bond). The three-dimensional networks of high atomic density consequently explained the occurrence of superhardness in ZrB4. The two structures are similar, but we find that Amm2-ZrB4 might be more easily obtained due to its lower formation enthalpy. These facts stimulate us to conduct a detailed investigation on its fundamental properties, such as elastic constants, elastic anisotropy and thermodynamic properties which are crucial to its practical applications and synthesis. For example, the elastic constants of a solid give important information concerning the nature of the forces operating in the solid13,14 and help us to understand its mechanical behaviors in practical application, such as anisotropy, phase transformation, elastic instability, plastic deformation and fracture, precipitation, dislocation dynamics, crack and so on.15 On the other hand, the understanding of thermodynamic properties of solids (such as heat capacity, thermal expansion coefficient, Grüneisen parameters, and Debye temperature) will be beneficial to their synthesis and practical applications.16 Therefore, in this paper, elastic properties of ZrB4 coupled with thermodynamic properties at various temperatures and pressures are investigated systematically through the first principles calculations and quasi-harmonic Debye model.17

This paper proceeds as follows: the details of the calculation methods and theoretical model are described in Section II, followed by the calculated results and analysis in Section III. Conclusions are summarized in Section IV.

II. Methods of calculation

The ab initio calculations were performed using density functional theory within the generalized gradient approximation (GGA),18 as implemented in the Vienna ab initio simulation package (VASP).19 The exchange and correlation potential was treated by the generalized gradient approximation in the scheme of Perdew–Burke–Ernzerhof (PBE).20 The all-electron projector augmented wave (PAW) method21 was employed with a plane-wave cutoff energy of 600 eV. The k-point grid in the Brillouin zone were generated using the Monkhorst–Pack scheme with the separation of 0.03 Å−1. The total energy convergence tests showed that convergence to within 1 meV/atom was achieved with the above calculation parameters. Single crystal elastic constants were calculated via a strain–stress approach. i.e., by applying a small strain to the equilibrium lattice of orthorhombic unit cell and fitting the dependence of the resulting change in stress on the strain. The bulk modulus, shear modulus, Young's modulus, and Poisson's ratio were determined by using the Voigt–Reuss–Hill approximation,22 in which the Voigt and Reuss expressions represent the upper and lower limit of the polycrystalline modulus. The formulae for orthorhombic structure are:
 
image file: c5ra15992j-t1.tif(1)

The Young's modulus E and the Poisson's ratio υ are then calculated from the elastic moduli using the following relations:

 
image file: c5ra15992j-t2.tif(2)

III. Results and discussion

3.1 Elastic properties

The lately predicted crystal structure of ZrB4 is orthorhombic with space group Amm2, no. 38, with 20 atoms per conventional unit cell consisting of four ZrB4 f.u. in a unit cell, in which one Zr and three B atoms occupy the Wyckoff position 8f (0.3461, 0.6689, 0.2655), 4c (0.8289, 0.5, 0.7469), 4e (0.5, 0.3374, 0.9792), and 4d (0, 0.3332, 0.2486) respectively. The equilibrium lattice constants, volume per formula unit, density, bulk modulus and its pressure derivative are all listed in Table 1, together with available experimental and theoretical results of ZrB2 and ZrB12 for comparison. In the elastic range, due to the symmetry of the crystal, there are nine independent components in the elastic tensor for ZrB4, i.e., C11, C22, C33, C44, C55, C66, C12, C13 and C23. Elastic constants play important roles in providing a deeper insight into mechanical stability and stiffness of materials.23 The pressure dependences of the elastic constants up to 100 GPa are illustrated in Fig. 1. It can be seen that all elastic constants increase monotonically with pressure and all Cij satisfy the well-known Born stability criteria24 up to 100 GPa, which indicates that ZrB4 is still mechanically stable at high pressure of 100 GPa.
 
image file: c5ra15992j-t3.tif(3)
Table 1 The calculated equilibrium lattice constants a0, b0, c0 (Å) and equilibrium volume per formula unit V03), density ρ, EOS fitted bulk modulus B0 (GPa), and its pressure derivative B0 for the orthorhombic ZrB4 at 0 K and 0 GPa
    a0 b0 c0 V0 ρ B0 B0
a Vinet universal EOS.26b Birch–Murnaghan 3rd-order EOS.27c Murnaghan EOS.28d Ref. 29.e Ref. 30.f Ref. 31.g Ref. 32.
ZrB4 This work 10.3120 5.41307 3.17999 300 5.03 239a, 238b, 235c 3.84a, 3.86b, 3.90c
ZrB2 Theo. 3.1768d 3.559d 31.1d 355d 4.2d
Exp. 3.170d 3.532d 30.74d 317f, 245g
ZrB12 Theo. 7.415e 407.69e
Exp. 7.4077e 406.49e



image file: c5ra15992j-f1.tif
Fig. 1 Pressure dependence of the elastic constants (Cij) of ZrB4 at 0 K.

Unfortunately, there are no experimental data available for comparison present, therefore, our results could be a reference for future studies and applications under high pressures of ZrB4.

In general, the large value of shear modulus is an indication of more pronounced directional bonding between atoms, and the Poisson's ratio is a factor that measures the stability of a crystal against shear and Young's modulus provides a measure of stiffness of a solid. The calculated elastic constants, elastic moduli (B, G and E), Poisson's ratio υ and the B/G ratio of ZrB4 under pressure are given in Table 2, along with the theoretical values of other transition metal tetraborides (WB4, CrB4, FeB4). It is shown in Table 2 that all of the B, G, E, υ and B/G increase substantially with pressure and the calculated bulk and shear moduli of Amm2-ZrB4 are comparable to those of WB4, CrB4, FeB4, indicating their strong ability to resist volume deformation. According to Pugh's criterion,25 a low (high) B/G value is associated with brittleness (ductility), and the ductile and brittle materials are separated by the critical value (1.75). The B/G of ZrB4 reaches 1.47 at 100 GPa, implying that ZrB4 is a brittle and mechanically stable phase within the range of pressures.

Table 2 The elastic constants Cij (GPa), bulk modulus B (GPa), shear modulus G (GPa), Young's modulus E (GPa), Poisson's ratio υ, the B/G ratio and the Hv of ZrB4 under pressure
  P C11 C22 C33 C44 C55 C66 C12 C13 C23 B G E ν B/G
a Ref. 10.b Ref. 3.
ZrB4 0 559 578 458 233 243 262 53 118 113 239 232 528 0.134 1.03
10 618 625 517 256 270 283 75 146 139 275 253 581 0.149 1.09
20 674 681 572 278 295 301 97 173 166 311 271 631 0.162 1.15
30 740 750 638 297 318 324 108 194 190 346 294 687 0.169 1.18
40 790 801 689 314 340 339 128 221 217 379 310 731 0.179 1.22
50 838 849 736 330 361 353 148 249 245 412 324 770 0.188 1.27
60 885 896 780 345 381 366 168 275 272 443 338 808 0.196 1.31
70 929 940 829 359 401 378 188 302 300 475 351 845 0.204 1.35
80 972 984 875 372 419 389 207 329 327 506 363 879 0.211 1.40
90 1016 1027 912 384 438 404 226 356 354 536 375 912 0.217 1.43
100 1057 1069 954 394 455 414 246 382 381 566 385 943 0.223 1.47
WB4a 0 389.3   437.0 150.7     280.2 224.2   297.0 103.6     2.86
CrB4b 0 554 880 473 254 282 250 65 107 95 265 261     1.02
FeB4b 0 381 710 435 218 114 227 137 143 128 253 177     1.43


Debye temperature θD is a fundamental parameter of a compound, which has close relationships with specific heat, melting temperature, and elastic constants. The θD can be calculated from elastic constants (image file: c5ra15992j-t4.tif),33 which gives explicit information about the lattice vibrations.34 The Debye temperature of ZrB4 under pressure is presented in Table 3, showing an increasing trend with pressure. As is generally known, a crystal with a larger Debye temperature corresponds to a stiffer characteristic. This is because the optical phonons have a higher frequency and therefore require greater energy to activate. Pressure typically enhances the interactions between atoms of a crystal and hence stiffers it, which is manifested by increased elastic moduli B and G. Therefore, pressure typically increases Debye temperature of ZrB4 and this implies stronger interactions between atoms in the system.

Table 3 The calculated density ρ (in g cm−3), the longitudinal, transverse and mean elastic wave velocity (υl, υt and υm in m s−1), and the Debye temperature θD (in K) of ZrB4 under pressure
P ρ υl υt υm θD
0 5.03 10[thin space (1/6-em)]457 6800 7457 1073
10 5.23 10[thin space (1/6-em)]821 6952 7634 1113
20 5.41 11[thin space (1/6-em)]150 7082 7788 1148
30 5.58 11[thin space (1/6-em)]499 7260 7989 1189
40 5.73 11[thin space (1/6-em)]757 7354 8100 1217
50 5.88 11[thin space (1/6-em)]983 7426 8188 1241
60 6.02 12[thin space (1/6-em)]188 7494 8270 1263
70 6.15 12[thin space (1/6-em)]385 7555 8344 1283
80 6.27 12[thin space (1/6-em)]565 7607 8407 1302
90 6.39 12[thin space (1/6-em)]731 7658 8469 1320
100 6.51 12[thin space (1/6-em)]883 7696 8517 1335


3.2 Elastic anisotropy

Elastic anisotropy is very important in diverse applications of materials, such as phase transformations, precipitation, dislocation dynamics and microcrack formation. The fundamental information about the bonding characteristics between adjacent atomic planes can also be obtained via the elastic anisotropy. Therefore, this property will be crucial for the potential hard material ZrB4. The shear anisotropy factors (A1, A2, A3), the universal elastic anisotropy index AU and the directional bulk modulus Ba, Bb and Bc are appropriate measures to quantify the extent of anisotropy.35 The shear anisotropy factor for the {100} shear planes between the <011> and <010> directions is defined as
 
image file: c5ra15992j-t5.tif(4)
for the {010} shear planes between the <101> and <001> directions is
 
image file: c5ra15992j-t6.tif(5)
for the {001} shear planes between the <110> and <010> directions is
 
image file: c5ra15992j-t7.tif(6)
for the universal elastic anisotropy index AU, defined by Ranganathan and Ostoja-Starzewski from the bulk modulus B and shear modulus G denoted by Voigt and Reuss approaches,35 is
 
image file: c5ra15992j-t8.tif(7)
and the directional bulk modulus along different crystallographic axis can be defined as36
 
Bi = i(dP/di) (i = a, b, and c) (8)

Taking advantage of the formulae mentioned above, the parameters about elastic anisotropy (A1, A2, A3, AU, Ba, Bb and Bc) are calculated and presented in Table 4. In the case of isotropic crystals, A1, A2, and A3 are all equal to 1, while any deviation from one means the amplitude of anisotropy of the crystal. From Table 4, we can see that A1, A2, and A3 are larger than 1 at 0 GPa and all increase with pressure. The shear anisotropy results of ZrB4 indicate that the elastic anisotropy for the {010} shear planes between the <101> and <001> directions is more obvious than that of the {100} shear planes between the <011> and <010> directions and the {001} shear planes between the <110> and <010> directions, and the value of A3 also reveals that ZrB4 is nearly isotropic in {001} shear planes. Because Amm2-ZrB4 is orthorhombic, the shear anisotropy factors are not adequate to sufficiently describe its elastic anisotropy. Therefore, the universal elastic anisotropy index AU should also be considered (AU is zero for isotropic crystals). In Table 4, AU is 0.09 at 0 GPa, which increases with increasing pressure. Meanwhile, the directional bulk modulus (Ba, Bb, Bc) also increases with pressure and the bulk modulus along the c-axis is larger than that along a-axis and b-axis at 100 GPa.

Table 4 The shear anisotropy factors A1, A2, A3 and elastic anisotropy index AU and the directional bulk modulus Ba, Bb and Bc of ZrB4 under pressure
P A1 A2 A3 AU Ba Bb Bc
0 1.194 1.173 1.001 0.090 742.4 786.1 655.8
10 1.215 1.249 1.035 0.098 853.7 853.5 775.58
20 1.234 1.280 1.038 0.099 958.3 956.7 885.98
30 1.199 1.260 1.017 0.091 1049.9 1059.1 1002.6
40 1.211 1.288 1.017 0.095 1144.3 1155.6 1112.1
50 1.224 1.319 1.017 0.101 1236.3 1249.5 1222.1
60 1.237 1.348 1.014 0.108 1325.7 1341.3 1322.2
70 1.243 1.369 1.012 0.112 1409.8 1427.7 1438.3
80 1.249 1.392 1.011 0.118 1492.0 1513.1 1553.2
90 1.261 1.425 1.015 0.131 1581.1 1600.7 1646.3
100 1.265 1.443 1.013 0.137 1660.3 1684.5 1755.1


Although the factors calculated above have already conveyed that the elastic properties of ZrB4 are anisotropic, it is still necessary to characterize the mechanical anisotropy in a more straightforward way. The shape of the 3D curved surface is sphere for isotropic materials (AU = 0), but for anisotropic materials, the sphere will deform. The degree of deformation reflects the extent of anisotropy, and the variation of elastic modulus with direction can be demonstrated. Therefore the Young's modulus, Shear modulus, and Poisson's ratio along different directions in three-dimensional (3D) space as well as the projections in (−110) plane and (001) plane at pressures 0 GPa, 50 GPa and 100 GPa have been drawn to denote the elastic anisotropy of ZrB4 on crystallographic directions, as is shown in Fig. 2 and 3. The direction dependent Young's modulus (E), Shear modulus (G) and Poisson's ratio (υ) for orthorhombic crystals36,37 can be defined respectively as:


image file: c5ra15992j-f2.tif
Fig. 2 Direction dependence of Young's modulus E (a), (d), (g), shear modulus G (b), (e), (h) and Poisson's ratio υ (c), (f), (i) under different pressures for ZrB4, the units are in GPa for E and G.

image file: c5ra15992j-f3.tif
Fig. 3 The projections of Young's modulus E (a), shear modulus G (b) and Poisson's ratio υ (c) in (−110) plane and (001) plane at pressures 0 GPa, 50 GPa and 100 GPa respectively, the units are in GPa for E and G.

Young's modulus:

 
E−1 = s11 = s11l14 + s22l24 + s33l34 + 2s12l12l22 + 2s23l22l32 + 2s13l12l32 + s44l22l32 + s55l12l32 + s66l12l22 (9)

Shear modulus:

 
G−1 = 4s11l12m12 + 4s22l22m22 + 4s33l32m32 + 8s12l1m1l2m2 + 8s23l2m2l3m3 + 8s13l1m1l3m3 + s44(l2m3 + m2l3)2 + s55(l1m3 + m1l3)2 + s66(l1m2 + m1l2)2 (10)

Poisson's ratio:

 
s12 = l12m12s11 + (l12m22 + l22m12)s12 + (l12m32 + l32m12)s13 + l22m22s22 + (l22m32 + l32m22)s23 + l32m32s33 + l2l3m2m3s44 + l1l3m1m3s55 + l1l2m1m2s66 (11)
 
image file: c5ra15992j-t9.tif(12)
where sij is the usual elastic compliance constants, li is the direction cosines in any arbitrary direction and mi is the direction cosines in perpendicular direction. From Fig. 2, elastic anisotropy is clearly seen in ZrB4, and the greater the pressure, the more obvious the anisotropy. In addition, the magnitude of Young's modulus in a specific direction can also be used to indicate the strength of chemical bonds in that direction. In Fig. 2(a), (d) and (g), the maximum of Young's modulus is observed in <111> direction, and the minimum occurs in <001> direction. Because a larger Young's modulus often stands for more covalent feature of a material,38,39 we can substantiate that the covalent feature of the bonding in <111> direction is more dominant than other directions. The G is remarkably dependent on the stress direction (Fig. 2(b), (e) and (h)) with the highest (lowest) value in the [001] ([111]) direction, and the Poisson's ratio (Fig. 2(c), (f) and (i)) has similar characteristics. Fig. 3(a) and (b) show the orientation dependence of E and G changing from [001] to [110] direction in (−110) plane and from [100] to [010] direction in (001) plane under different pressures, and the shape of the projections in (001) plane at 0 GPa, 50 GPa, 100 GPa is almost round, which illustrates that ZrB4 is nearly isotropic in (001) plane. This result is consistent with the shear anisotropy factor A3. Poisson's ratio represents the negative ratio of transverse and longitudinal strains which plays a significant role in mechanical engineering design.40 The values of υ in (−110) plane varies in a very large range as shown in Fig. 3(c), the features of υ under 100 GPa are 0.283 < υ < 0.381 in (−110) plane. It means that when the stress direction is perpendicular to the (−110) plane, the maximum strain is noted in the [001] direction and the minimum strain in the [111] direction.

3.3 Thermodynamic properties

The thermodynamic properties of ZrB4 at various temperatures (0–1000 K) and pressures (0–50 GPa) are systematically calculated. In Fig. 4, we present the normalized volume–pressure and bulk modulus–pressure diagram of ZrB4 at temperatures 0, 200, 400, 600, 800, and 1000 K, where V0 is the zero-pressure equilibrium volume. It is easily seen from Fig. 4 that, as pressure increases, the relative volume V/V0 decreases at a given temperature and the V/V0 curve becomes steeper with temperature increasing, which implies that ZrB4 is more easily compressed when temperature increases. Furthermore, it is found that the bulk modulus increases with pressure at a constant temperature and decreases with temperature at a given pressure.
image file: c5ra15992j-f4.tif
Fig. 4 The calculated normalized volume V/V0 and bulk modulus of ZrB4 as a function of pressure at temperatures 0, 200, 400, 600, 800, and 1000 K.

The calculated heat capacity of ZrB4 as a function of temperature (pressure) at given pressure (temperature) is demonstrated in Fig. 5. It is shown in Fig. 5(a) that the heat capacity CV fits T3 term in their sufficiently low-temperature regions and approximates to absolute zero when the temperature vanishes at the given 0, 10, 20, 30, 40, 50 GPa. This is due to the harmonic approximations of the Debye model used here. At intermediate temperatures, the temperature dependence of CV is dominated by the details of vibrations of atoms.41 At high temperatures, the calculated CV is expected to get close to the Dulong–Petit limit, 3nNAkB (n is the number of atoms in a molecule, NA is the Avogadro constant and kB is the Boltzmann constant), which is common to all solids at high temperatures. For ZrB4, the Dulong–Petit limit is about 120 J mol−1 K−1. The pressure dependence of the heat capacity for ZrB4 at 100, 200, 400, 600, 800 and 1000 K is presented in Fig. 5(b). It is noted that the calculated CV decreases with pressure at a constant temperature and increases with temperature at a given pressure. From Fig. 5(a) and (b), we can get that the effect of temperature on CV is greater than that of pressure.


image file: c5ra15992j-f5.tif
Fig. 5 (a) Temperature dependence of heat capacity at different pressures and (b) pressure dependence of heat capacity at various temperatures.

The volume thermal expansion coefficient α as a function of temperature (pressure) at different pressures (temperatures) is shown in Fig. 6(a) and (b). Because of the weak dependence of the bulk modulus on temperature and that α is proportional to CV (image file: c5ra15992j-t10.tif, γ is the Gruneisen parameter, and K is the bulk modulus), the trend of the volume thermal expansion coefficient is similar to the heat capacity. As shown in Fig. 6(a), at given pressures, α increases rapidly with temperature at sufficiently low temperatures (α(T) ∼ T3) and gradually turns to a slow increase at high temperatures (T > 400 K). Additionally, it is noted in Fig. 6(b) that α decreases with increasing pressure at a constant temperature, and the trend slows down at high pressures.


image file: c5ra15992j-f6.tif
Fig. 6 (a) Temperature dependence of thermal expansion coefficient at different pressures. (b) Pressure dependence of the thermal expansion coefficient at various temperatures.

Fig. 7 shows the pressure dependence of the Debye temperature θD and Grüneisen parameter γ of ZrB4 at different temperatures (0, 100, 200, 400, 600, 800, and 1000 K). It is easily seen in Fig. 7(a) that when temperature keeps constant, Debye temperature increases almost linearly with increasing pressure and compared with pressure, the variation of θD caused by temperature is very small. Therefore, we can draw a conclusion that the effect of the temperature on θD is not as significant as that of pressure. In quasi-harmonic Debye model, Grüneisen parameter γ describes the anharmonic effects of the crystal lattice thermal vibration. From Fig. 7(b), we can see that at fixed temperature, γ decreases sharply with pressure, and as temperature goes higher, γ decreases more rapidly with the increase of pressure.


image file: c5ra15992j-f7.tif
Fig. 7 Debye temperature θD (a) and Grüneisen parameter γ (b) for ZrB4 as a function of pressure at different temperatures.

IV. Conclusion

In conclusion, we have focused our attention on prediction and detailed analysis of elastic constants, anisotropic properties, and thermodynamic properties under high pressures of Amm2-ZrB4 by first principles calculations in combination with the quasi-harmonic Debye model in this work. In the light of the Born stability criteria and the Pugh criterion, ZrB4 (Amm2) is mechanically stable and exhibits brittle nature within the scope of the studied pressure (0–100 GPa). The Debye temperature of ZrB4 was calculated by taking advantage of the relationship that Θ is proportional to the averaged sound velocity vm, and it increases with pressure. Young's modulus, shear modulus and Poisson's ratio as a function of crystal orientation have been systematically investigated and analyzed. ZrB4 exhibits pronounced elastic anisotropy and the extent increases with pressure. Furthermore, the pressure and temperature dependences of calculated normalized volume V/V0, bulk modulus, volume thermal expansion coefficient, specific heat, Debye temperature, and Grüneisen parameter have also been evaluated in the ranges of 0–50 GPa and 0–1000 K through quasi-harmonic Debye model. The results point out that pressure and temperature have manifest effects on these thermodynamic properties. The present study provides detailed and systematic information for Amm2-ZrB4, which is of fundamental importance for its industrial application.

Acknowledgements

This work was supported by the NBRPC (grant 2013CB733000), NSFC (grants 51171160/51434008/11447243). S. Z. would like to acknowledge support by the Natural Science Foundation of Hebei province, China (grant E2015203200).

References

  1. Q. Gu, G. Krauss and W. Steurer, Adv. Mater., 2008, 20, 3620 CrossRef CAS PubMed .
  2. A. R. Oganov, J. Chen, C. Gatti, Y. Ma, Y. Ma, C. W. Glass, Z. Liu, T. Yu, O. O. Kurakevych and V. L. Solozhenko, Nature, 2009, 457, 863 CrossRef CAS PubMed .
  3. H. Niu, J. Wang, X.-Q. Chen, D. Li, Y. Li, P. Lazar, R. Podloucky and A. N. Kolmogorov, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 144116 CrossRef .
  4. J. Haines, J. Leger and G. Bocquillon, Annu. Rev. Mater. Res., 2001, 31, 1 CrossRef CAS .
  5. X. Zhang, J. Qin, J. Ning, X. Sun, X. Li, M. Ma and R. Liu, J. Appl. Phys., 2013, 114, 183517 CrossRef PubMed .
  6. X. Zhang, J. Qin, Y. Xue, S. Zhang, Q. Jing, M. Ma and R. Liu, Phys. Status Solidi RRL, 2013, 7, 1022 CrossRef CAS PubMed .
  7. H. Gou, N. Dubrovinskaia, E. Bykova, A. A. Tsirlin, D. Kasinathan, W. Schnelle, A. Richter, M. Merlini, M. Hanfland and A. M. Abakumov, Phys. Rev. Lett., 2013, 111, 157002 CrossRef .
  8. Y. Chen, D. He, J. Qin, Z. Kou, S. Wang and J. Wang, J. Mater. Res., 2010, 25, 637 CrossRef CAS .
  9. C. Liu, F. Peng, N. Tan, J. Liu, F. Li, J. Qin, J. Wang, Q. Wang and D. He, High Pressure Res., 2011, 31, 275 CrossRef CAS PubMed .
  10. M. Wang, Y. Li, T. Cui, Y. Ma and G. Zou, Appl. Phys. Lett., 2008, 93, 101905 CrossRef PubMed .
  11. T. Tokunaga, K. Terashima, H. Ohtani and M. Hasebe, Mater. Trans., 2008, 49, 2534 CrossRef CAS .
  12. X. Zhang, J. Qin, X. Sun, Y. Xue, M. Ma and R. Liu, Phys. Chem. Chem. Phys., 2013, 15, 20894 RSC .
  13. N. Korozlu, K. Colakoglu, E. Deligoz and S. Aydin, J. Alloys Compd., 2013, 157 CrossRef CAS PubMed .
  14. S. Zhang, X. Zhang, Y. Zhu, S. Zhang, L. Qi and R. Liu, Intermetallics, 2014, 44, 31 CrossRef CAS PubMed .
  15. X. Zhang, J. Qin, T. Perasinjaroen, W. Aeksen, M. K. Das, R. Hao, B. Zhang, P. Wangyao, Y. Boonyongmaneerat, S. Limpanart, M. Ma and R. Liu, Surf. Coat. Technol., 2015, 276, 228 CrossRef CAS PubMed .
  16. R. Hao, X. Zhang, J. Qin, S. Zhang, J. Ning, N. Sun, M. Ma and R. Liu, RSC Adv., 2015, 5, 36779 RSC .
  17. M. Blanco, E. Francisco and V. Luana, Comput. Phys. Commun., 2004, 158, 57 CrossRef CAS PubMed .
  18. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS .
  19. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS .
  20. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS .
  21. G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758 CrossRef CAS .
  22. R. Hill, Proc. Phys. Soc., 1952, 65, 349 CrossRef .
  23. S. Zhang, X. Zhang, Y. Zhu, M. Ma, J. Qin and R. Liu, Mater. Chem. Phys., 2015, 149–150, 553 CrossRef CAS PubMed .
  24. M. Born, Proc. Cambridge Philos. Soc., 1940, 36, 160 CrossRef CAS .
  25. S. F. Pugh, Philos. Mag., 1954, 45, 823 CrossRef CAS PubMed .
  26. P. Vinet, J. H. Rose, J. Ferrante and J. R. Smith, J. Phys.: Condens. Matter, 1989, 1, 1941 CrossRef CAS .
  27. F. Birch, Phys. Rev., 1947, 71, 809 CrossRef CAS .
  28. F. D. Murnaghan, Am. J. Appl. Math., 1937, 235 Search PubMed .
  29. H. Fu, M. Teng, X. Hong, Y. Lu and T. Gao, Phys. B, 2010, 405, 846 CrossRef CAS PubMed .
  30. A. Rybina, K. Nemkovski, P. Alekseev, J.-M. Mignot, E. Clementyev, M. Johnson, L. Capogna, A. Dukhnenko, A. Lyashenko and V. Filippov, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 024302 CrossRef .
  31. A. Pereira, C. Perottoni, J. da Jornada, J. Leger and J. Haines, J. Phys.: Condens. Matter, 2002, 14, 10615 CrossRef CAS .
  32. N. L. Okamoto, M. Kusakari, K. Tanaka, H. Inui, M. Yamaguchi and S. Otani, J. Appl. Phys., 2003, 93, 88 CrossRef CAS PubMed .
  33. O. L. Anderson, J. Phys. Chem. Solids, 1963, 24, 909 CrossRef CAS .
  34. J. Jia, D. Zhou, J. Zhang, F. Zhang, Z. Lu and C. Pu, Comput. Mater. Sci., 2014, 95, 228 CrossRef CAS PubMed .
  35. S. I. Ranganathan and M. Ostoja-Starzewski, Phys. Rev. Lett., 2008, 101, 055504 CrossRef .
  36. P. Ravindran, L. Fast, P. Korzhavyi, B. Johansson, J. Wills and O. Eriksson, J. Appl. Phys., 1998, 84, 4891 CrossRef CAS PubMed .
  37. J. F. Nye, Physical Properties of Crystals: Their Representation by Tensors and Matrices, Oxford University Press, Oxford, 1985 Search PubMed .
  38. M. Rajagopalan, S. P. Kumar and R. Anuthama, Phys. B, 2010, 405, 1817 CrossRef CAS PubMed .
  39. G. Yi, X. Zhang, J. Qin, J. Ning, S. Zhang, M. Ma and R. Liu, J. Alloys Compd., 2015, 640, 455 CrossRef CAS PubMed .
  40. J. Lewandowski, W. Wang and A. Greer, Philos. Mag. Lett., 2005, 85, 77 CrossRef CAS PubMed .
  41. Z. Huang, J. Feng and W. Pan, Comput. Mater. Sci., 2011, 50, 3056 CrossRef CAS PubMed .

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.