Abstract
By utilizing the first-principles method, the pressure-induced effects on phase transition, mechanical stability, and elastic properties of WS2/CrS2 are investigated in the pressure range from 0 to 80 GPa. Transitions from 2Hc to 2Ha for WS2 and CrS2 are found to occur at 17.5 and 25 GPa, respectively. It is found that both 2Ha and 2Hc phases of WS2 and CrS2 meet the mechanical stability criteria up to 80 GPa, suggesting that those structures are mechanically stable. The bulk and shear modulus anisotropy of the two phases of WS2 and CrS2 decrease rapidly under pressure and, finally, trend to isotropy. With increasing pressure, the elastic moduli (Y, B, and G), sound velocities (vs, vp, vm), and Debye temperatures (Θ) of 2Ha and 2Hc of WS2 and CrS2 increase monotonously. Moreover, the Debye temperature (Θ) of 2Hc phase is higher than that of 2Ha phase for both WS2 and CrS2. The bulk, shear, and Young’s modulus, Poisson coefficient, and brittle/ductile behaviour are estimated. The percentages of anisotropy in compressibility and shear and the ratio of bulk to shear modulus (B/G) are also studied.
1 Introduction
There is a growing interest in studding the structures and properties of two-dimensional transition metal dichalcogenides (TMDs) owing to their novel electronic and catalytic properties that differ from their bulk counterparts [1–3]. Because TMDs have in-plane covalent bonding and weak interlayer van der Waals interactions, they could be easily exfoliated down to a monolayer which shows very exotic properties. For example, the band-gap structure of MoS2 shows an indirect-to-direct semiconductor transition from bulk to single-layer due to a lack of interlayer interaction [4, 5]. Two-dimensional dilute magnetic semiconductors are proposed for substitution of Mo by other transition metal atoms, such as W and Cr. Furthermore, TMDs also have shown exciting prospects for various applications, such as lubricants and catalysts in the petroleum industry [6], possible applications in optoelectronics and nanoelectronics [7], and energy-storage applications [8]. As an important member of TMDs, WS2 has received a lot of attention. For example, WS2 has been prepared into nanoparticles, nanotubes, nanoribbons, films, graphene-like WS2, and WS2 sheets [9–16]. Some features and properties of WS2 such as tribological performance [10, 17, 18], structural, electronic [19], optical, and mechanical properties [4, 9, 20–24] have been studied. CrS2 also attracted some researchers’ interest, and the electronic properties [25, 26], optical, piezoelectric properties, and stability of this single-layer material were investigated and characterized [27].
However, most of the previous theoretical works focused on WS2 and CrS2 were carried out at 0 GPa. Very recently, high pressure in situ angle dispersive powder X-ray diffraction using synchrotron radiation together with high-pressure Raman analysis was performed on powder WS2 samples to investigate the effect of pressure on crystal WS2. However, it was reported that there was no phase transformations observed in WS2 under compression for pressures up to 52 GPa [28]. As we have known, for most transition typical TMDs, there is a phase transition from 2Hc to 2Ha under pressure [29, 30]. As we know, due to some reasons, such as the limitations of the experimental conditions, temperature effect, and higher phase transition barrier, some phase transition under pressure cannot be detected experimentally. So it is necessary to investigate structural evolution of WS2 theoretically. Since chromium belongs to the same group as tungsten, we also investigate the possible pressure-induced transition from 2Hc to 2Ha for CrS2.
In this paper, we first investigate the possible structural transitions of WS2/CrS2 under pressure, then study the anisotropy in elasticity and thermodynamic properties of them under pressure. We found that both CrS2 and WS2 will transform from 2Hc to 2Ha phase under pressure, and the anisotropy in elasticity and thermodynamic properties of 2Hc and 2Ha phases for both WS2/CrS2 are calculated and compared.
The paper is structured as follows. Section 2 contains the computational details. The results and discussion are provided in Section 3. Finally, we present our conclusions in Section 4.
2 Computational Methods
The calculations reported in this work were performed using the CASTEP code in the Materials Studio software [31, 32]. The density functional theory of the plane-wave pseudopotential method [33] was used to perform geometry optimisation and calculate the elastic parameters. The exchange and correlation functional were treated by the local density approximation [34, 35]. The plane-wave cutoff energy of 800 eV was used for the all structures studied in this work. The k integration over the Brillouin zone was performed up to a 14×14×4 Monkhorst–Pack [36] mesh. Coulomb potential energy caused by electron–ion interaction is described using Norm-Conserving Pseudopotential [37]. In order to make our calculations standard, energy-difference threshold for the convergence is taken as 5*10–6 eV/atom. These values ensure a satisfactory convergence, and thus provide reliable results.
3 Results and Discussion
The two typical structures of TMDs (2Hc and 2Ha) [29] are shown in Figure 1, and the pressure-induced sliding of layers was reported to be responsible for the transition from 2Hc to 2Ha structure [29]. The calculated lattice constants of 2Hc structures for both CrS2 and WS2 at zero pressure together with other theoretical values are listed in Table 1, and our results agree well with other reference’s data. These guarantee the reliability of our calculation.

(a) Structures of 2Hc-XS2 (X=W, Cr). (b) Structures of 2Ha-XS2 (X=W, Cr).
Calculated lattice constant a (Å) at zero pressure, together with this work and other references for comparison.
Reference | Lattice constant a/Å | Lattice constant c/Å | |
---|---|---|---|
WS2 | This work | 3.1462 | 12.163 |
[38] | 3.1532 | 12.323 | |
[39] | 3.154 | 12.362 | |
CrS2 | This work | 3.04 | 14.409 |
[40] | 3.06 | ||
[41] | 3.05 |
To investigate the possible pressure-induced structural transition of CrS2 and WS2, we applied hydrostatic pressure to 2Hc and 2Ha phases, and then fully optimised the lattice parameters and atomic positions. Figure 2a and b plot the relative enthalpy versus pressure for the 2Hc and 2Ha structures of CrS2 and WS2, respectively. For CrS2, 2Ha structure is with lower enthalpy than 2Hc structure for pressures ≥17.5 GPa. For WS2, enthalpy of 2Ha phase is lower than that of 2Hc phase at 25 GPa. So, for CrS2 and WS2, the phase-transition pressures for 2Hc to 2Ha phase are 17.5 and 25 GPa, respectively.

(a) Enthalpy versus pressure for the 2Hc and 2Ha structures of WS2. (b) Enthalpy versus pressure for the 2Hc and 2Ha structures of CrS2.
The elastic properties of a material are very important for understanding and predicting the mechanical stability, and for indications of material-related properties such as brittleness, ductility, based on the analysis of elastic constants Cij, bulk modulus B, and shear modulus G. The elastic constants can be obtained by calculating the stress response when a small strain was forced to the optimised unit cell [42]. According to the strain energy theory, for a mechanically stable structural phase, the strain energy should be positive, and the matrix of elastic constants should be positive, definite, and symmetrical [43]. For hexagonal structure, the mechanical stability criterion can be expressed as [44]:
According to the calculated elastic constants of the hexagonal 2Hc and 2Hc phase of CrS2 and WS2, it is found that they all satisfy their stability conditions up to 80 GPa. That means, for CrS2 and WS2, both the 2Ha and 2Hc hexagonal phases are mechanically stable up to 80 GPa. In the view of mechanical stability criteria, 2Hc phase may exist as metastable structure at higher pressure.
The changes of the elastic constants with pressure for both CrS2 and WS2 are presented in Figures 3 and 4, respectively. It can be found that the elastic constants C11 and C33 increase quickly and monotonically with pressure, while the elastic constants C44, C12, and C13 increase comparatively slowly. For these two materials, C11 and C33 are larger than other elastic constants, which indicate that they are very incompressible under uniaxial stress along x (ε11) or z (ε33) axis.

(a) Elastic constants of 2Ha phase of CrS2 as a function of pressure. (b) Elastic constants of 2Hc-phase of CrS2 as a function of pressure.

(a) Elastic constants of 2Hc phase of WS2 as a function of pressure. (b) The elastic constants of 2Ha phase of WS2 as a function of pressure.
Bulk modulus reflects the compressibility of the solid under hydrostatic pressure. The bulk modulus (B) and shear modulus (G) can be estimated according to the Voigt–Reuss–Hill approximation [45–48]. For the hexagonal systems, they can be calculated with the computed data using the following relations:
The elastic anisotropy of crystals has a great impact on the properties of physical mechanism, such as anisotropic plastic deformation, elastic instability, and crack behavior. Hence, it is important to calculate the elastic anisotropy to improve its mechanical durability. The universal anisotropic index (AU) is represented as
The percentage elastic anisotropy in compressibility and shear can be used as follows:
where B and G denote the bulk and shear modulus, and the subscripts V and R represent the Voigt and Reuss bounds. For these two formulas, a value of zero is associated with isotropic elastic constants, while a value of 1 (100 %) is the largest possible anisotropy. The calculated anisotropic index and percent anisotropy factors are shown in Tables 2 and 3, respectively. It can be seen that the two phases of WS2 and CrS2 are anisotropic under the ordinary condition. With increasing pressure, the calculated universal anisotropic index (AU) decreases. So, both the anisotropies of 2Ha- and 2Hc-WS2 and CrS2 decrease with pressure. Furthermore, the results also indicate that the 2Ha phase of WS2 and CrS2 has stronger mechanical anisotropy than the 2Hc phase. The value of bulk modulus anisotropy and shear modulus anisotropy decreases first, and they will be close to zero under certain pressure (such as the value of AB of 2Ha-WS2 under 70 GPa). So, as the pressure increases, the bulk modules and shear modulus of two structural of WS2 and CrS2 become slightly anisotropic and, finally, turn to isotropic under certain pressures.
Anisotropic factors, Debye temperature, average sound velocity of 2Hc, and 2Ha-WS2.
Pressure (Gpa) | WS2 (2Ha phase) | WS2 (2Hc phase) | ||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
B | G | AU | AB | AG | Vm | Θ | B | G | AU | AB | AG | Vm | Θ | |
0 | 54 | 37 | 9.111 | 0.4519 | 0.4273 | 2428 | 277 | 60 | 48 | 4.382 | 0.3105 | 0.2583 | 2729 | 313 |
10 | 130 | 81 | 0.815 | 0.1009 | 0.0558 | 3353 | 401 | 133 | 91 | 0.547 | 0.0275 | 0.0468 | 3557 | 425 |
20 | 179 | 108 | 0.438 | 0.0544 | 0.0313 | 3743 | 458 | 181 | 120 | 0.174 | 0.0048 | 0.0162 | 3950 | 482 |
30 | 220 | 135 | 0.407 | 0.0278 | 0.0339 | 4062 | 506 | 220 | 145 | 0.062 | 0.0004 | 0.0061 | 4233 | 525 |
40 | 263 | 152 | 0.518 | 0.0133 | 0.0468 | 4235 | 535 | 256 | 168 | 0.021 | 0.0000 | 0.0022 | 4463 | 561 |
50 | 303 | 173 | 0.521 | 0.0034 | 0.0489 | 4430 | 566 | 271 | 189 | 0.010 | 0.0000 | 0.0010 | 4633 | 590 |
60 | 341 | 189 | 0.590 | 0.0007 | 0.0557 | 4565 | 590 | 324 | 213 | 0.027 | 0.0036 | 0.0020 | 4847 | 624 |
70 | 376 | 204 | 0.679 | 0.0000 | 0.0637 | 4676 | 610 | 361 | 232 | 0.043 | 0.0065 | 0.0030 | 4991 | 649 |
80 | 412 | 215 | 0.790 | 0.0003 | 0.0733 | 4750 | 625 | 397 | 250 | 0.057 | 0.0085 | 0.0039 | 5112 | 671 |
Anisotropic factors, Debye temperature, average sound velocity of 2Hc, and 2Ha-CrS2.
Pressure (GPa) | CrS2 (2Ha phase) | CrS2 (2Hc phase) | ||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
B | G | AU | AB | AG | Vm | Θ | B | G | AU | AB | AG | Vm | Θ | |
0 | 52 | 34 | 6.168 | 0.3898 | 0.3284 | 3782 | 448 | 58 | 44 | 2.879 | 0.2663 | 0.1772 | 4258 | 508 |
10 | 99 | 51 | 7.419 | 0.3973 | 0.3789 | 4326 | 540 | 126 | 83 | 0.261 | 0.0256 | 0.0204 | 5487 | 682 |
20 | 178 | 107 | 0.354 | 0.0262 | 0.0292 | 5962 | 764 | 166 | 106 | 0.060 | 0.0095 | 0.0041 | 5997 | 763 |
30 | 214 | 131 | 0.327 | 0.0103 | 0.0297 | 6429 | 838 | 195 | 134 | 0.013 | 0.0029 | 0.0008 | 6542 | 848 |
40 | 254 | 146 | 0.455 | 0.0051 | 0.0426 | 6663 | 881 | 235 | 162 | 0.019 | 0.0003 | 0.0019 | 7022 | 925 |
50 | 295 | 155 | 0.688 | 0.0008 | 0.0643 | 6747 | 903 | 272 | 186 | 0.044 | 0.0022 | 0.0040 | 7372 | 984 |
60 | 332 | 174 | 0.581 | 0.0000 | 0.0549 | 7037 | 952 | 307 | 208 | 0.076 | 0.0054 | 0.0065 | 7669 | 1035 |
70 | 369 | 182 | 0.836 | 0.0005 | 0.0771 | 7113 | 971 | 341 | 231 | 0.096 | 0.0070 | 0.0081 | 7953 | 1085 |
80 | 404 | 195 | 0.848 | 0.0016 | 0.0779 | 7267 | 1001 | 379 | 253 | 0.120 | 0.0082 | 0.0103 | 8210 | 1130 |
Debye temperature is closely related to many physical properties of solids, such as specific heat, stability of lattices, and melting temperature. At low temperature, it can be calculated using the average sound velocity (vm) according to the following [49]:
where h is the Plank’s constant, k is the Boltzmann’s constant, and V is the atomic volume. The average sound velocity (vm) in the polycrystalline material can be obtained using the following [49]:
where νs and νp are the transverse and longitudinal sound velocity, and are given by Navier’s equation [49]:
Depending upon the above equations, the average sound velocity and Debye temperature of 2Hc- and 2Ha-WS2/CrS2 are calculated and listed in Tables 2 and 3, respectively. It can be seen that vm and Θ of 2Hc- and 2Ha-WS2/CrS2 increase with increasing pressure. Additionally, the calculated results predict that Θ of 2Hc phase is higher than that of 2Ha phase for both WS2 and CrS2.
Knowing G and B, the Young’s modulus Y and Poisson’s ratio δ, which are frequently measured for polycrystalline materials when investigating their hardness, can be calculated from the isotropic relations:
The pressure dependencies of Young’s modulus Y and shear modulus G of CrS2 and WS2 at various pressures are plotted in Figure 5a and b, respectively. It can be seen that both 2Hc and 2Ha phases of CrS2 and WS2 show similar trends under pressure. The values of G and Y increase with increasing pressure, and both of them were affected significantly by the pressure. Furthermore, it is noted from Figure 5 that G and Y change approximately linearly with the pressure, and the values of G and Y of 2Hc phase are larger than that of 2Ha phase for both CrS2 and WS2.

(a) Pressure dependencies of the shear modulus G and Young’s modulus Y of CrS2. (b) Pressure dependencies of the shear modulus G and Young’s modulus Y of WS2.
Poisson’s ratio δ is related to the volume change during elastic deformation [50]. It can be used to assess the stability of a crystal against shear. The range of values for Poisson’s ratio is –1 to 0.5, with –1 indicating the material does not change its shape and 0.5 indicating the volume remains unchanged during elastic deformation. The obtained Poisson’s ratio for 2Hc- and 2Ha-CrS2 at pressure 0–80 GPa are 0.194–0.227 and 0.23–0.29, respectively, which indicate that both 2Hc- and 2Ha-CrS2 have central interatomic forces and are relatively stable against shear. As shown in Figure 6a, the Poisson’s ratio δ of 2Hc- and 2Ha-CrS2 increases with increasing pressure, and 2Ha phase has larger δ than 2Hc phase. The changes of δ with pressure for 2Hc- and 2Ha-WS2 are shown in Figure 6b, and it can be seen that the change trends of δ are similar with that of CrS2.

(a) Pressure dependencies of the Poisson’s ratio and the ratio of bulk to shear modulus (B/G) of CrS2. (b) Pressure dependencies of the Poisson’s ratio and the ratio of bulk to shear modulus (B/G) of WS2.
Considering that the bulk modulus B represents the resistance to fracture, while the shear modulus G represents the resistance to plastic deformation, the ratio of bulk to shear modulus (B/G) was introduced to measure the empirical malleability of polycrystalline materials [51]. The value at the critical point which separates ductile and brittle materials is about 1.75. If B/G>1.75, a material behaves in a ductile manner, and if not, a material will demonstrate brittleness. Higher (lower) values of B/G mean higher ductile (brittle) of a material. The B/G ratios of the two phases of CrS2 at pressure from 0 to 80 GPa are shown in Figure 6a, and it is found that with increasing pressure, the B/G value of 2Hc phase increases from 1.30 to 1.49, while the B/G value of 2Ha phase increases from 1.53 to 2.07. Therefore, the 2Hc phase of WS2 is brittle over the pressure range of 0–80 GPa, while 2Ha phase changes from brittle to ductile with increasing pressure. Similar features are also found in the B/G ratios of WS2. The B/G ratios of WS2 of the two phases at pressure from 0 to 80 GPa are shown in Figure 6b; with increasing pressure, the B/G value of 2Hc phase increases from 1.24 to 1.58, while the B/G value of 2Ha phase increases from 1.45 to 1.91. Therefore, the 2Hc phase of WS2 is brittle up to 80 GPa, while the 2Ha phase changes from brittle to ductile with pressure increasing.
4 Conclusions
In this work, the pressure-induced phase transition, mechanical stability, and elastic properties of 2Hc- and 2Ha-WS2/CrS2 are investigated using the first-principles method. The calculated equilibrium lattice parameters of 2Hc-WS2/CrS2 at 0 GPa are well in agreement with the experimental and other theoretical values. Although the transition for 2Hc to 2Ha phase was not observed experimentally for WS2, we found that CrS2 and WS2 will transform from 2Hc to 2Ha phase at 17.5 and 25 GPa, respectively. According to the mechanical stability criterion, both 2Hc- and 2Ha-WS2/CrS2 exhibit mechanical stability in the pressure range of 0–80 GPa. The shear modulus, Young’s modulus, Debye temperature, and average sound velocity of 2Hc- and 2Ha-WS2/CrS2 increase with the increasing pressure. The changes of Poisson’s ratio and ratio of bulk to shear modulus (B/G) are also studied, and our results show that the 2Hc phase of WS2/CrS2 is brittle on the whole pressure range we investigated, while the 2Ha phase changes from brittle to ductile with pressure increasing. We also found that the two phases of WS2 and CrS2 are anisotropic under the ordinary condition, while the mechanical anisotropy of 2Ha- and 2Hc-WS2 and CrS2 decreases as pressure increases.
Acknowledgments:
This work was supported by the Natural Science Foundation of China (Grant Nos. U1304612, U1404608, and 51501093), Young Core Instructor Foundation of Henan Province (Grant No. 2015GGJS-122), Science Technology Innovation Talents in Universities of Henan Province (Grant No. 16HASTIT047), Innovation Scientists and Technicians Troop Construction Projects of Henan Province (Grant No. C20150029), and Scientific Research Fund of Henan Provincial Education Department (Grant No. 15B140006). Feiwu Zhang acknowledges the support of the Thousand Young Talents Program and the Hundred Talent Program of the Chinese Academy of Sciences.
References
[1] Y. Li, Z. Zhou, S. Zhang, and Z. Chen, J. Am. Chem. Soc. 130, 16739 (2008).10.1021/ja805545xSearch in Google Scholar
[2] K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, Phys. Rev. Lett. 105, 136805 (2010).10.1103/PhysRevLett.105.136805Search in Google Scholar
[3] A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, et al., Nano Lett. 10, 1271 (2010).10.1021/nl903868wSearch in Google Scholar
[4] Y. Ding, Y. Wang, J. Ni, L. Shi, S. Shi, et al., Phys. B 406, 2254 (2011).10.1016/j.physb.2011.03.044Search in Google Scholar
[5] S. Lebègue and O. Eriksson, Phys. Rev. B 79, 5409 (2009).10.1103/PhysRevB.79.115409Search in Google Scholar
[6] P. Raybaud, J. Hafner, G. Kresse, and H. Toulhoat, J. Phys. Condens. Matter 9, 11107 (1997).10.1088/0953-8984/9/50/014Search in Google Scholar
[7] Q. Wang, K. Kalantar-Zadeh, A. Kis, J. Coleman, and M. Strano, Nat. Nanotech. 7, 699 (2012).10.1038/nnano.2012.193Search in Google Scholar
[8] M. Chhowalla, H. S. Shin, G. Eda, J. L. Li, K. P. Loh, et al., Nat. Chem. 5, 263 (2013).10.1038/nchem.1589Search in Google Scholar
[9] A. Kumar and P. K. Ahluwalia, Eur. Phys. J. B 85, 186 (2012).10.1140/epjb/e2012-30070-xSearch in Google Scholar
[10] C. Drummond, N. Alcantar, J. Israelachvili, R. Tenne, and Y. Golan, Adv. Funct. Mater. 11, 348 (2001).10.1002/1616-3028(200110)11:5<348::AID-ADFM348>3.0.CO;2-SSearch in Google Scholar
[11] W. Ho, J. C. Yu, J. Lin, J. G. Yu, and P. Li, Langmuir 20, 5865 (2004).10.1021/la049838gSearch in Google Scholar
[12] A. K. M. Newaz, D. Prasai, J. I. Ziegler, D. Caudel, S. Robinson, et al., Solid State Commun. 155, 49 (2013).10.1016/j.ssc.2012.11.010Search in Google Scholar
[13] K. K. Kam and B. A. Parklnson, J. Phys. Chem. 86, 463 (1982).10.1021/j100393a010Search in Google Scholar
[14] Z. Wang, K. Zhao, H. Li, Z. Liu, Z. Shi, et al., J. Mater. Chem. 21, 171 (2011).10.1039/C0JM02821ESearch in Google Scholar
[15] H. S. S. Ramakrishna Matte, A. Gomathi, A. K. Manna, D. J. Late, R. Datta, et al., Angew. Chem. 122, 4153 (2010).10.1002/ange.201000009Search in Google Scholar
[16] M. Nath, A. Govindaraj, and C. N. R. Rao, Adv. Mater. 13, 283 (2001).10.1002/1521-4095(200102)13:4<283::AID-ADMA283>3.0.CO;2-HSearch in Google Scholar
[17] S. V. Prasad and J. S. Zabinski, J. Mater. Sci. Lett. 12, 1413 (1993).10.1007/BF00591592Search in Google Scholar
[18] L. F. Mattheis, Phys. Rev. Lett. 30, 784 (1973).10.1103/PhysRevLett.30.784Search in Google Scholar
[19] G. Arora, Y. Sharma, V. Sharma, G. Ahmed, S. K. Srivastava, et al., J. Alloys Compd. 470, 452 (2009).10.1016/j.jallcom.2008.02.098Search in Google Scholar
[20] I. Kaplan-Ashiri and R. Tenne, J. Clus. Sci. 18, 549 (2007).10.1007/s10876-007-0118-9Search in Google Scholar
[21] A. Kuc, N. Zibouche, and T. Heine, Phys. Rev. B 83, 245213 (2011).10.1103/PhysRevB.83.245213Search in Google Scholar
[22] H. Jiang, J. Phys. Chem. C 116, 7664 (2012).10.1021/jp300079dSearch in Google Scholar
[23] T. Björkman, A. Gulans, A. V. Krasheninnikov, and R. M. Nieminen, Phys. Rev. Lett. 108, 235502 (2012).10.1103/PhysRevLett.108.235502Search in Google Scholar PubMed
[24] L. P. Feng, Z. Q. Wang, and Z. T. Liu, Solid State Commun. 187, 43 (2014).10.1016/j.ssc.2014.02.012Search in Google Scholar
[25] S. Lebègue, T. Björkman, M. Klintenberg, R. M. Nieminen, and O. Eriksson, Phys. Rev. X 3, 031002 (2013).10.1103/PhysRevX.3.031002Search in Google Scholar
[26] C. Ataca, H. Sahin, and S. Ciraci, J. Phys. Chem. C 116, 8983 (2012).10.1021/jp212558pSearch in Google Scholar
[27] H. L. Zhuang, M. D. Johannes, M. N. Blonsky, and R. G. Hennig, Appl. Phys. Lett. 104, 022116 (2014).10.1063/1.4861659Search in Google Scholar
[28] N. Bandaru, R. S. Kumar, J. Baker, O. Tschauner, T. Hartmann, et al., Int. J. Mod. Phys. B 28, 134 (2014).10.1142/S0217979214501689Search in Google Scholar
[29] L. Hromadová, R. Martoňák, and E. Tosatti, Phys. Rev. B 87, 144105 (2013).10.1103/PhysRevB.87.144105Search in Google Scholar
[30] N. Bandaru, R. S. Kumar, D. Sneed, O. Tschauner, J. Baker, et al., J. Phys. Chem. C 118, 3230 (2014).10.1021/jp410167kSearch in Google Scholar
[31] S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. J. Probert, et al., Z. Kristallogr. 220, 567 (2005).10.1524/zkri.220.5.567.65075Search in Google Scholar
[32] M. D. Segall, P. J. D. Lindan, M. J. Probert, C. J. Pickard, P. J. Hasnip, et al., J. Phys. Condens. Matter 14, 2717 (2002).10.1088/0953-8984/14/11/301Search in Google Scholar
[33] P. Hohenberg and W. Kohn, Phys. Rev. B 136, 864 (1964).10.1103/PhysRev.136.B864Search in Google Scholar
[34] D. M. Ceperley and B. J. Alder, Phys. Rev. Lett. 45, 566 (1980).10.1103/PhysRevLett.45.566Search in Google Scholar
[35] J. P. Perdew and A. Zunger, Phys. Rev. B 23, 5048 (1981).10.1103/PhysRevB.23.5048Search in Google Scholar
[36] H. J. Monkhorst and J. D. Pack, Phys. Rev. B 13, 5188 (1976).10.1103/PhysRevB.13.5188Search in Google Scholar
[37] D. R. Hamann, M. Schluter, and C. Chiang, Phys. Rev. Lett. 43, 1494 (1979).10.1103/PhysRevLett.43.1494Search in Google Scholar
[38] C. Lee, J. Hong, M. H. Whangbo, and J. H. Shim, Chem. Mater. 25, 3745 (2013).10.1021/cm402281nSearch in Google Scholar
[39] A. Matthäus, A. Ennaoui, S. Fiechter, S. Tiefenbacher, T. Kiesewetter, et al., J. Electrochem. Soc. 144, 1013 (1997).10.1149/1.1837522Search in Google Scholar
[40] X. L. Wei, H. Zhang, G. C. Guo, X. B. Li, W. M. Lau, et al., J. Mater. Chem. A 2, 2101 (2014).10.1039/C3TA13659KSearch in Google Scholar
[41] H. Guo, N. Lu, L. Wang, X. Wu, and X. C. Zeng, J. Phys. Chem. C 118, 7242 (2014).10.1021/jp501734sSearch in Google Scholar
[42] B. B. Karki, L. Stixrude, S. J. Clark, M. C. Warren, G. J. Ackland, et al., Am. Mineral. 82, 51 (1997).10.2138/am-1997-1-207Search in Google Scholar
[43] M. H. Sadd, Elasticity: Theory, Applications, and Numerics, 2nd ed., Academic Press, Burlington, MA 2009.Search in Google Scholar
[44] G. V. Sin’Ko and N. A. Smirnov, J. Phys.: Condens. Matter. 14, 6989 (2002).10.1088/0953-8984/14/29/301Search in Google Scholar
[45] W. Voigt, Lehrbuch der kristallphysik: Teubner-Leipzig, Macmillan, New York 1928.Search in Google Scholar
[46] A. Reuss and Z. Angew, Math. Mech. 9, 55 (1929).10.1002/zamm.19290090104Search in Google Scholar
[47] R. Hill, Proc. Phys. Soc. London, Sect. A 65, 349 (1952).10.1088/0370-1298/65/5/307Search in Google Scholar
[48] X. H. Deng, W. Lu, Y. M. Hu, and H. S. Gu, Phys. B 404, 1218 (2009).10.1016/j.physb.2008.11.200Search in Google Scholar
[49] O. L. Anderson, J. Phys. Chem. Solids 24, 909 (1963).10.1016/0022-3697(63)90067-2Search in Google Scholar
[50] P. Ravindran, L. Fast, P. A. Korzhavyi, B. Johansson, J. Wills, et al., J. Appl. Phys. 84, 4891 (1998).10.1063/1.368733Search in Google Scholar
[51] S. F. Pugh, Philos. Mag. 45, 823 (1954).10.1080/14786440808520496Search in Google Scholar
©2016 by De Gruyter
Articles in the same Issue
- Frontmatter
- Spinless Particle in a Magnetic Field Under Minimal Length Scenario
- Constraint on the Multi-Component CKP Hierarchy and Recursion Operators
- Band Structure Characteristics of Nacreous Composite Materials with Various Defects
- The Integrability of an Extended Fifth-Order KdV Equation in 2+1 Dimensions: Painlevé Property, Lax Pair, Conservation Laws, and Soliton Interactions
- Numerical Solution for the Effect of Suction or Injection on Flow of Nanofluids Past a Stretching Sheet
- First-Principles Calculations of the Mechanical and Elastic Properties of 2Hc- and 2Ha-WS2/CrS2 Under Pressure
- Conservation Laws and Mixed-Type Vector Solitons for the 3-Coupled Variable-Coefficient Nonlinear Schrödinger Equations in Inhomogeneous Multicomponent Optical Fibre
- Classical Equation of State for Dilute Relativistic Plasma
- Magnetic Field and Slip Effects on the Flow and Heat Transfer of Stagnation Point Jeffrey Fluid over Deformable Surfaces
- Nonlocal Symmetry and its Applications in Perturbed mKdV Equation
- Universality of the Phonon–Roton Spectrum in Liquids and Superfluidity of 4He
Articles in the same Issue
- Frontmatter
- Spinless Particle in a Magnetic Field Under Minimal Length Scenario
- Constraint on the Multi-Component CKP Hierarchy and Recursion Operators
- Band Structure Characteristics of Nacreous Composite Materials with Various Defects
- The Integrability of an Extended Fifth-Order KdV Equation in 2+1 Dimensions: Painlevé Property, Lax Pair, Conservation Laws, and Soliton Interactions
- Numerical Solution for the Effect of Suction or Injection on Flow of Nanofluids Past a Stretching Sheet
- First-Principles Calculations of the Mechanical and Elastic Properties of 2Hc- and 2Ha-WS2/CrS2 Under Pressure
- Conservation Laws and Mixed-Type Vector Solitons for the 3-Coupled Variable-Coefficient Nonlinear Schrödinger Equations in Inhomogeneous Multicomponent Optical Fibre
- Classical Equation of State for Dilute Relativistic Plasma
- Magnetic Field and Slip Effects on the Flow and Heat Transfer of Stagnation Point Jeffrey Fluid over Deformable Surfaces
- Nonlocal Symmetry and its Applications in Perturbed mKdV Equation
- Universality of the Phonon–Roton Spectrum in Liquids and Superfluidity of 4He