Abstract
The structural parameters, electronic structures, and mechanical and thermodynamic properties of TaSi2 under different pressures have been completely explored by a combination of density functional theory and quasi-harmonic Debye model. Results show that our computed structural parameters and elastic constants are in consistency with available experimental findings and previous theoretical calculations. The electronic structures of TaSi2 under different pressures including band structures and density of states are reported. It turns out that TaSi2 should be metallic. The elastic constants Cij, bulk modulus B, shear modulus G, Young’s modulus E, Poisson’s ratio ν, B/G, Debye temperature θ, and wave velocities under pressures are also evaluated successfully. The calculated Cij obeys the Born–Huang stability criterion, which demonstrates that TaSi2 is mechanically stable under different pressures. More interestingly, the three-dimensional surface constructions and projections of E and B under different pressures are also systematically evaluated. With the increase of applied pressure, TaSi2 exhibits subtle anisotropy under zero pressure, and the anisotropy strengthened. Finally, the dependence of the thermodynamic properties on pressure/temperature is obtained and analyzed for the first time.
1 Introduction
Binary transition metal disilicides have been widely investigated because of their applications in the fields of heterostructure devices and thermoelectric materials [1], [2], [3]. Many transition metal disilicides are crystallizing in C40, C11b, and C54 structures [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15]. NbSi2, VSi2, and TaSi2 are C40 structures (space group: P6222) [7], [8], [9], [10], [11], [12], [13], [14], [15]. MoSi2 and WSi2 are tetragonal structures (space group: I4/mmm) [4], [5]. TiSi2 crystallizes in the C54 structure (space group Fddd

Crystal structure of TaSi2.
The optimized and experimental lattice constants a and c (Å), c/a ratio, and volume V (Å3) of TaSi2.
Pressure (GPa) | Source | a | c | c/a ratio | V |
---|---|---|---|---|---|
0 | This work | 4.80 | 6.60 | 1.375 | 131.69 |
Exp. (Ref. [7]) | 4.77 | 6.55 | 1.373 | 129.07 | |
Exp. (Ref.[11]) | 4.784 | 6.568 | 1.373 | 130.18 | |
10 | This work | 4.7212 | 6.5125 | 1.379 | 125.72 |
20 | This work | 4.6538 | 6.4420 | 1.384 | 120.82 |
30 | This work | 4.5965 | 6.3808 | 1.388 | 116.75 |
40 | This work | 4.5467 | 6.3265 | 1.391 | 113.26 |
50 | This work | 4.5024 | 6.2780 | 1.394 | 110.21 |
2 Computational Methods
Our theoretical calculations on TaSi2 were performed with DFT as implemented in the Vienna Ab Initio Simulation Package [16], [17]. The projector augmented wave (PAW) [18] method for pseudopotentials was performed to describe the electron-core interaction. For Ta atom, 5s, 5p, 5d, and 6s orbitals and, for Si atom, 3s and 3p orbitals are treated as valence states. The exchange-correlation functionals was employed using the generalized gradient approximation with the Perdew-Burke-Ernzerhof [19]. A plane wave energy cut-off of 480 eV was used throughout our calculations. The irreducible Brillouin Zone was sampled by an 8 × 8 × 8 grid, determined according to the Monkhorst-Pack k-point meshes. The total energy changes during the optimization finally converged to 10−6 eV/cell, and the Hellman–Feymann force per atom was reduced to 0.02 eV/Å. All above parameter settings have been made to obtain good convergence.
3 Results
3.1 Structural Parameters
The hexagonal structure (space group of P6222) of TaSi2 was displayed in Figure 1. In this hexagonal structure, Ta and Si atoms are located in the Wyckoff 3d (0.5, 0, 0.5) and 6j (0.1593, 0.3185, 0.5) sites, respectively.

Band structures of TaSi2 at (a) 0 and (b) 50 GPa.
By performing PAW calculations, we obtained the lattice constants for a = 4.80 Å, c = 6.60 Å, c/a = 1.375, and volume V = 131.69 Å3 (Table 1), which is well consistent with experimental findings: a = 4.784 Å, c = 6.568 Å, c/a = 1.373, V = 130.18 Å3, as well as previous experimental findings [7]. This indicates the accuracy of our PAW calculations.
3.2 Electronic Properties
Using the optimized lattice constants, a systematic study on band structures and density of states were carried out using the generalized gradient approximation (Perdew–Burke–Ernzerhof) functional at pressure P = 0 and 50 GPa, respectively. Representative graphs of the band structures are plotted in Figure 2 with the Fermi level (EF) set to zero. A shift of position under pressure at P = 0 and 50 GPa can be observed in Figure 2. We observe that some bands cross the EF, confirming that TaSi2 is metallic. The total densities of states together with partial densities of states of TaSi2 at P = 0 GPa are plotted in Figure 3. Figure 3 shows that the band structure of TaSi2 can be divided into two regions. The conduction bands around 0–5 eV for TaSi2 are dominated by Ta d orbital, whereas the valence bands from −15 to 0 eV is mainly Ta d orbital and Si s hybridized with the Si p orbital.

Total and partial densities of states of TaSi2 at 0 GPa.
The calculated elastic constants Cij (GPa) and elastic compliance matrix Sij (10−3 GPa−1) of TaSi2 under different pressures.
Pressure (GPa) | C11 | C12 | C13 | C33 | C44 | S11 | S12 | S13 | S33 | S44 |
---|---|---|---|---|---|---|---|---|---|---|
0 | 368.82 | 80.18 | 93.90 | 457.19 | 138.50 | 2.951 | −0.514 | −0.500 | 2.393 | 7.220 |
Ref. [15] | 351 | 84 | 73 | 461 | 123 | |||||
Ref. [7] (room temperature) | 375.73 | 78.4 | 90.1 | 476.7 | 143.7 | 2.875 | −0.493 | −0.450 | 2.268 | 7.006 |
10 | 434.49 | 98.73 | 109.29 | 526.92 | 150.93 | 2.514 | −0.464 | −0.425 | 2.074 | 6.626 |
20 | 492.8 | 117.16 | 123.58 | 591.41 | 160.72 | 2.227 | −0.436 | −0.374 | 1.847 | 6.222 |
30 | 548.52 | 135.51 | 137.42 | 651.01 | 168.81 | 2.009 | −0.412 | −0.337 | 1.678 | 5.924 |
40 | 601.57 | 154.15 | 149.93 | 708.49 | 175.64 | 1.840 | −0.395 | −0.306 | 1.541 | 5.693 |
50 | 652.14 | 172.91 | 161.46 | 764.49 | 181.54 | 1.704 | −0.383 | −0.279 | 1.426 | 5.508 |
3.3 Elastic Properties
The obtained single crystal elastic constants Cij (C11, C12, C13, C33, and C44) under various pressures (0–50 GPa) are shown in Table 2, as well as the independent constants of elastic compliance matrix calculated from elastic constants. Additionally, to further confirm the mechanical stability of the hexagonal structure, we predicted the mechanical stability via the Born–Huang stability criterion. For the hexagonal crystal system, the calculated elastic constants need to satisfy the following stability conditions [20]:

The evolution of calculated elastic constants of TaSi2 as a function of pressure.
From Table 2, we find that the calculated elastic constants satisfy the above conditions. As mentioned above, it is indicated that the hexagonal structure is a mechanically stable structure. The calculated elastic constants at zero temperature and under some fixed pressures compared with available previous experimental findings (at room temperature) are presented in Table 2, being in consistency with the experimental findings (at room temperature) [7] and other theoretical results [15]. We also explored the pressure evolution of elastic constants of TaSi2 (Fig. 4). We can find that the elastic constants increase with the applied pressure. With the increase of the pressure from 0 to 50 GPa, the C11, C12, C13, C33, and C44 of TaSi2 at zero temperature is increased by 43.44 %, 53.63 %, 41.84 %, 40.20 %, and 23.71 %, respectively.
On the basis of the known Cij values, the bulk modulus B and shear modulus G can be computed by the Voigt approximation [21]:
and the Reuss approximation, which happens to be formally equivalent to the single crystal bulk modulus in this case [22]:

The calculated B, G, and E of TaSi2 as a function of pressure.
The calculated Bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), B/G, Poisson’s ratio ν, wave velocities Vp, VS, and Vm (km/s), and Debye temperatures θ (K) of TaSi2 under different pressures.
Pressure (GPa) | B | G | E | B/G | ν | Vp | VS | Vm | θ |
---|---|---|---|---|---|---|---|---|---|
0 | 191.06 | 145.59 | 348.29 | 1.312 | 0.196 | 6.557 | 4.031 | 4.449 | 545.54 |
Ref. [15] | 179 | 137 | 327 | 1.307 | 0.195 | 526 | |||
Ref. [7] (room temperature) | 192.5 | 151.0 | 359.0 | 1.275 | 0.1892 | 552 | |||
10 | 224.58 | 165.08 | 397.78 | 1.360 | 0.205 | 6.880 | 4.192 | 4.630 | 583.33 |
20 | 255.24 | 181.45 | 440.06 | 1.407 | 0.213 | 7.131 | 4.308 | 4.763 | 614.26 |
30 | 284.57 | 196.15 | 478.50 | 1.451 | 0.220 | 7.347 | 4.403 | 4.871 | 641.04 |
40 | 312.55 | 209.56 | 513.85 | 1.491 | 0.226 | 7.534 | 4.483 | 4.963 | 664.80 |
50 | 339.38 | 221.92 | 546.61 | 1.529 | 0.232 | 7.699 | 4.550 | 5.041 | 686.10 |
Therefore, shear modulus upper bounded by can be presented as [21]
and lower bounded by [22]
Voigt and Reuss limits can be averaged, as proposed by Hill [23], to produce a single estimation of the bulk modulus and shear modulus of the TaSi2:
BV, BR, GV, and GR are the bulk moduli and shear moduli referring to the Voigt approximations and Ruess approximations, respectively.
The B, G, and E under various pressures (0–50 GPa) are illustrated in Figure 5. It can be clearly seen that calculated B, G, and E of TaSi2 increase with the applied pressure. In Table 3, we also listed the values of the B, G, and E with the pressure of TaSi2. The B, G, and E at 0 K and 0 GPa are 191.06, 145.59, and 348.29 GPa, which are in accord with experimental results (at room temperature) of 192.5, 151, and 359.0 GPa, respectively [7]. The Poisson’s ratio ν is 0.196, which is almost equal to the experimental value of 0.1892 [7]. On the basis of the Pugh rule (B/G = 1.75) [24], the ductility or brittleness of a solid is deduced by the B/G ratio. If B/G > 1.75, the solid’s behavior is ductile. On the contrary, the solid is brittle. From Table 3, it is found that TaSi2 under different pressures (0–50 GPa) show a brittle behavior because the obtained B/G ratio of TaSi2 is smaller than 1.75. In this work, the microhardness of C40 TaSi2 is calculated by the following five empirical rules [25], [26]:
We obtain that the maximum value is 29.5 GPa, and the minimum is 18.4 GPa. The results from the above rules tend to overestimate the experimental value (15.6 GPa) [27].
The anisotropy in elasticity can be characterized by universal anisotropic index (AU), the percent anisotropy (AB and AG), and shear anisotropic factors (A1, A2, and A3), which can be expressed from the following equations [25], [28], [29]:
The calculated universal anisotropic index (AU), percent anisotropy (AG and AB), and shear anisotropic factors (A1, A2, and A3) of TaSi2 under different pressures.
Pressure (GPa) | AU | AB (%) | AG (%) | A1 | A2 | A3 |
---|---|---|---|---|---|---|
0 | 0.045 | 0.657 | 0.320 | 0.868 | 0.868 | 1 |
10 | 0.056 | 0.460 | 0.468 | 0.855 | 0.855 | 1 |
20 | 0.077 | 0.369 | 0.690 | 0.813 | 0.813 | 1 |
30 | 0.102 | 0.297 | 0.947 | 0.768 | 0.768 | 1 |
40 | 0.131 | 0.239 | 1.244 | 0.730 | 0.730 | 1 |
50 | 0.163 | 0.198 | 1.569 | 0.695 | 0.695 | 1 |

The 3D Young’s modulus E curved surface of TaSi2 under (a) 0 and (c) 50 GPa. The 3D bulk modulus B curved surface under (b) 0 and (d) 50 GPa.
From Table 4, it is noted that percent anisotropy AG is larger than AB under different pressures (0–50 GPa) of TaSi2. It means that the shear modulus shows stronger directional dependence than the bulk modulus. The results of A1, A2, and A3 under different pressures (0–50 GPa) are also illustrated in Table 4. If A1 = A2 = A3 = 1, the material is an isotropic crystal. Otherwise, it is an anisotropic crystal. The results indicate that it is an anisotropic crystal. The deviations of universal anisotropic index (AU) under different pressures (0–50 GPa) from zero mean the highly single crystal anisotropy.

The projections in different planes of Young’s modulus E of TaSi2. (a) XY plane, (b) XZ plane and YZ plane. The units are in GPa.

The projections in different planes of bulk modulus B of TaSi2. (a) XY plane, (b) XZ plane and YZ plane. The units are in GPa.
In this work, the anisotropy of the three-dimensional (3D) surface of E and B of TaSi2 under 0 and 50 GPa is presented in Figure 6. As seen in Figure 6, the nonspherical shape shows the extent of the elastic anisotropy, especially, and the anisotropy behavior becomes more obvious with the increasing pressure. In Figures 7 and 8, we, respectively plotted the projections of B and E at 0, 10, 20, 30, 40, and 50 GPa. As shown in Figures 7 and 8, it is found that the projections of XZ and YZ planes are the same. From Figures 7 and 8, the projections of E on the XY, XZ, and YZ planes show more anisotropic behavior than B. From the projections of E and B, it is concluded that anisotropic behavior of TaSi2 strengthened with increasing pressure.

The calculated wave velocities (a) and Debye temperature (b) of TaSi2 as a function of pressure.
As we know, the Debye temperature θ can be observed from the average wave velocity [30], [31]. The calculated θ as well as the compressional velocity VP, the shear wave velocity VS, and the average wave velocity Vm under different pressures (0–50 GPa) are shown in Table 3 and Figure 9. As seen in Figure 9, the wave velocities and θ increase gradually with applied pressure. As shown in Table 3, at T = 0 and P = 0, θ is 545.54 K, which is a bit smaller than the experimental value of 552 K of Chu [7] and slightly more than the theoretical value of 526 K [15].
The melting temperatures (Tm) of hexagonal crystals are related to their elastic constants C11 and C33 [32]. The melting temperature (Tm) is defined as follows:
where C11 and C33 is in GPa and Tm in K. We obtain that the Tm of TaSi2 is 2146.2 K, which is slightly less than the measured Tm of 2473.15 K for TaSi2 [7].

The calculated and fitted V/V0 as a function of pressure.
3.4 Thermodynamic Properties
In this work, the thermodynamic properties were evaluated using the quasiharmonic Debye model as implemented in the Gibbs code [33]. We obtained the thermodynamic properties under different temperatures (0–2100 K) and pressures (0–30 GPa).
The dependence of the relative volume V/V0 on temperature/pressure is plotted in Figures 10 and 11. As shown in Figure 10, we found that present fitted values are consistent with calculated data. From Figure 11, it should be noted that V/V0 decreases with increasing pressure. It is also obtained that V/V0 tends to decrease with increasing temperature at a given pressure.
In Figure 12, we pictured the relationships between the B and pressure/temperature. Figure 12a displays the calculated B with pressure at different temperatures (0, 500, 1000, and 2000 K). One can see that the calculated B increases linearly with pressure at a given temperature. From Figure 12b, the magnitude of B decreases linearly with increasing temperature at different pressures (0, 10, 20, and 30 GPa). It is worth noticing that P has a significant effect on B than T. At zero pressure and T = 300 K, our calculated B value (187.54 GPa) satisfactorily agrees with the experimental value of 192.5 GPa (at room temperature) [7].

Contours of relative volume V/V0 versus pressure P (GPa) and temperature T (K).

The calculated B of TaSi2 as a function of (a) pressure and (b) temperature.
From Figure 13a, we can see at different temperatures (500, 1000, 1500, and 2000 K) that the calculated thermal expansion coefficient α decreases dramatically with increasing pressure. Figure 13b shows thermal expansion coefficient a of TaSi2 at different pressures (0, 10, 20, 30 GPa), from which it can be seen that the thermal expansion coefficient α increases quickly at a given pressure particularly below the temperature of 500 K. After a sharp increase, the thermal expansion coefficient of the TaSi2 is nearly insensitive to the temperature above 500 K because of the electronic contributions. The thermal expansion coefficient α decreases strongly with increasing pressure at a constant temperature. Our present calculated a is 0.76 × 10−5 K−1 at P = 0 and T = 300 K and decreases by 21.74 %, 36.30 %, and 46.60 % with pressure increasing to 10, 20, and 30 GPa, respectively.

The calculated α of TaSi2 as a function of (a) temperature and (b) pressure.

The calculated (a) CV and (b) CP of TaSi2 at different pressures as a function of temperature.
The heat capacity at constant volume CV and the heat capacity at constant pressure CP of TaSi2 as a function of temperature at some given pressure (0, 10, and 30 GPa) are plotted in Figure 14. The data of CV and CP are almost identical at low temperature, and the CV and CP of TaSi2 under different pressures (0, 10, and 30 GPa) are proportional to T3 at low temperature. At high temperature, CV is following the Dulong–Petit limit. CP increases monotonically with temperature at high temperature. It is concluded that the influence of T on the CV and CP are more significant than P of TaSi2. We can see that the variation of Cp as a function of temperature is similar to α. The influence of P on α is more significant than CP. The calculated CP is equal to 0.00265 J mol−1 K−1 at 5 K, which is larger than the experimental value of 0.000914 J mol−1 K−1 [9].
The calculated U (kJ mol−1), A (kJ mol−1), S (J mol−1 K−1), CV (J mol−1 K−1), CP (J mol−1 K−1), a (10−5 K−1), γ, and θ (K) of TaSi2 at 300 K under different pressures.
Pressure (GPa) | U | A | S | CV | CP | a | γ | θ |
---|---|---|---|---|---|---|---|---|
0 | 27.43 | 12.4 | 50.12 | 59.87 | 60.14 | 0.765 | 1.909 | 649.02 |
5 | 27.92 | 13.75 | 47.21 | 58.58 | 58.80 | 0.6705 | 1.853 | 681.62 |
10 | 28.35 | 14.90 | 44.86 | 57.45 | 57.63 | 0.599 | 1.802 | 709.87 |
15 | 28.79 | 15.99 | 42.68 | 56.32 | 56.48 | 0.5369 | 1.753 | 737.58 |
20 | 29.20 | 16.94 | 40.84 | 55.31 | 55.45 | 0.4879 | 1.710 | 762.25 |
25 | 29.6 | 17.86 | 39.13 | 54.31 | 54.43 | 0.444 | 1.668 | 786.46 |
30 | 29.98 | 18.68 | 37.65 | 53.40 | 53.51 | 0.408 | 1.631 | 808.38 |
Furthermore, we have also completely explored the internal energy U, entropy S, Helmholtz free energy A, CV, CP, a, Grüneisen parameter γ, and θ of TaSi2 at 300 K under different pressures in Table 5. So far as we have known, there are no experimental and theoretical findings available for comparison. So, our thermodynamic data provide an interesting insight and feasible guidelines for the potential applications of TaSi2.
4 Conclusions
We have utilized DFT and quasi-harmonic Debye model to investigate the structural parameters, electronic structures, mechanical property, and thermodynamic properties of TaSi2. The lattice constants and elastic constants computed agree well with experimental findings. The electronic structures under different pressures indicate that TaSi2 has a metallic behavior. The graphs of 3D and projections of E and B show that TaSi2 has an elastic anisotropy. Moreover, the thermal physical properties on pressure/temperature have been obtained and analyzed in detail for the first time. Our present theoretical predictions will be useful for experimental study on these transition-metal disilicides.
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 51502259
Funding source: “Six Top Talents” Program of Jiangsu Province
Award Identifier / Grant number: 2016-XCL-070
Funding source: Ministry of Housing and Urban-Rural Development of China
Award Identifier / Grant number: 2015-K4-007
Funding statement: This work was supported by the National Natural Science Foundation of China (51502259), the “Six Top Talents” Program of Jiangsu Province (2016-XCL-070), the Science and Technology Project from the Ministry of Housing and Urban-Rural Development of China (2015-K4-007), and the Top-Notch Academic Programs Project of Jiangsu Higher Education Institutions (no. PPZY2015A025).
References
[1] K. N. Tu and J. W. Mayer, in: Thin Films Interdiffusion and Interactions (Eds. J. M. Poate, K. N. Tu, J. W. Mayer), Wiley, New York 1978.Search in Google Scholar
[2] U. Gottlieb, F. Nava, M. Affronte, O. Laborde, and R. Madar, in: Properties of Metal Disilicides (Eds. K. Maex, M. van Rossum), EMIS Data Review Series No. 14, INSPEC, the Institution of Electrical Engineers, London 1995, p. 189.Search in Google Scholar
[3] P. A. Badoz, E. Rosencher, and J. Torres, J. Appl. Phys. 62, 890 (1987).10.1063/1.339695Search in Google Scholar
[4] X. L. Liu, Y. Ren, H. Xu, and Z. W. Zhao, J. Cent. South Univ. Technol. 17, 888 (2010).10.1007/s11771-010-0572-7Search in Google Scholar
[5] G. L. Xu, D. L. Zhang, Y. Z. Xia, X. F. Liu, Y. F. Liu, et al., Chin. Phys. Lett. 26, 046302 (2009).10.1088/0256-307X/26/4/046302Search in Google Scholar
[6] P. Ravindran, L. Fast, P. A. Korzhavyi, and B. Johansson, J. Appl. Phys. 84, 4891 (1998).10.1063/1.368733Search in Google Scholar
[7] F. Chu, M. Lei, S. A. Maloy, J. J. Petrovic, and T. E. Mitchell, Acta Mater. 44, 3035 (1996).10.1016/1359-6454(95)00442-4Search in Google Scholar
[8] F. Chu, M. Lei, S. A. Maloy, T. E. Mitchell, A. Miglori, and J. Garrett, Philos. Mag. 71, 373 (1995).10.1080/13642819508239040Search in Google Scholar
[9] O. Labordea, J. Bossy, M. Affronte, H. Schober, and U. Gottlie, Solid State Commun. 126, 415 (2003).10.1016/S0038-1098(03)00085-1Search in Google Scholar
[10] O. P. Balkashin, A. G. M. Jansen, U. Gottlieb, O. Laborde, and R. Madar, Solid State Commun. 100, 293 (1996).10.1016/0038-1098(96)00478-4Search in Google Scholar
[11] J. C. Lasjaunias, O. Laborde, and U. Gottlieb, J. Low Temp. Phys. 92, 335 (1993).10.1007/BF00682295Search in Google Scholar
[12] N. Xu, Y. Xu, and J. Ma, Mat. Sci. Semicon. Proc. 30, 636 (2015).10.1016/j.mssp.2014.09.030Search in Google Scholar
[13] H. J. Hou, H. J. Zhu, W. H. Cheng, and L. H. Xie, Z. Naturforsch. A 71, 657 (2016).10.1515/zna-2015-0367Search in Google Scholar
[14] E. Ertürk and T. Gürel, Physica B Condens Matter 537, 188 (2018).10.1016/j.physb.2018.01.070Search in Google Scholar
[15] B. Wan, F. R. Xiao, Y. K. Zhang, Y. Zhao, L. L. Wu, J. W. Zhang, and H. Y. Gou, J. Alloys Comp. 681, 412 (2016).10.1016/j.jallcom.2016.04.253Search in Google Scholar
[16] G. Kresse and J. Furthmuller, Comput. Mat. Sci. 6, 15 (1996).10.1016/0927-0256(96)00008-0Search in Google Scholar
[17] G. Kresse and J. Furthmuller, Phys. Rev. B 54, 11169 (1996).10.1103/PhysRevB.54.11169Search in Google Scholar PubMed
[18] G. Kresse and J. Joubert, Phys. Rev. B 59, 1758 (1999).10.1103/PhysRevB.59.1758Search in Google Scholar
[19] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).10.1103/PhysRevLett.77.3865Search in Google Scholar PubMed
[20] G. V. Sin’ko and N. A. Smirnow, J. Phys. Condens. Matter. 14, 6989 (2002).10.1088/0953-8984/14/29/301Search in Google Scholar
[21] W. Voigt, Lehrbuch der Kristallphysik, Terubner, Leipzig 1928.Search in Google Scholar
[22] A. Z. Reuss, Angew. Math. Mech. 9, 49 (1929).10.1002/zamm.19290090104Search in Google Scholar
[23] R. Hill, Proc. Phys. Soc. Sect. A 65, 349 (1952).10.1088/0370-1298/65/5/307Search in Google Scholar
[24] S. F. Pugh, Philos. Mag. 45, 823 (1954).10.1080/14786440808520496Search in Google Scholar
[25] A. L. Ivanovskii, Int. J. Refract. Met. Hard Mater. 36, 179 (2013).10.1016/j.ijrmhm.2012.08.013Search in Google Scholar
[26] L. Nie, W. Zhou, and Y. Zhan, Comput. Theor. Chem. 1020, 51 (2013).10.1016/j.comptc.2013.07.035Search in Google Scholar
[27] G. Schultes, M. Schmitt, D. Goettel, and O. Freitag-Weber, Sen. Actuators A 126, 287 (2006).10.1016/j.sna.2005.05.023Search in Google Scholar
[28] A. Guechi, A. Merabet, M. Chegaar, A. Bouhemadou, and N. Guechi, J. Alloys Compd. 623, 219 (2015).10.1016/j.jallcom.2014.10.114Search in Google Scholar
[29] S. I. Ranganathan and M. Ostoja-Starzewski, Phys. Rev. Lett. 101, 55504 (2008).10.1103/PhysRevLett.101.055504Search in Google Scholar
[30] K. B. Panda and K. S. Ravi Chandran, Comput. Mater. Sci. 35, 134 (2006).10.1016/j.commatsci.2005.03.012Search in Google Scholar
[31] O. L. Anderson, J. Phys. Chem. Solids 24, 909 (1963).10.1016/0022-3697(63)90067-2Search in Google Scholar
[32] M. E. Fine, L. D. Brown, and H. L. Marcus, Scripta Metall. 18, 951 (1984).10.1016/0036-9748(84)90267-9Search in Google Scholar
[33] M. A. Blanco, E. Francisco, and V. Luana, Comput. Phys. Commun. 158, 57 (2004).10.1016/j.comphy.2003.12.001Search in Google Scholar
©2019 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- General
- Analytical Solutions and Integrable Structure of the Time-Dependent Harmonic Oscillator With Friction
- Atomic, Molecular & Chemical Physics
- Wetting Behaviour of Gold Electrode and Molten Alkali Chlorides
- Probing Resonances and Pseudospin Symmetry of the Eckart Potential by the Complex Scaling Method within the Relativistic Framework
- Dynamical Systems & Nonlinear Phenomena
- First integrals and exact solutions of some compartmental disease models
- Gravitation & Cosmology
- Motion of a Test Particle According to the Scalar Ether Theory of Gravitation and Application to its Celestial Mechanics
- Hydrodynamics
- Analysis for Various Effects of Relaxation Time and Wall Properties on Compressible Maxwellian Peristaltic Slip Flow
- Acoustic Instability with Dynamic Charging in Quantum Plasmas
- Solid State Physics & Materials Science
- Preparation, Structural, Spectroscopic, Thermal, Linear and Nonlinear Optical Characteristics of Semi-Organic Material: Samarium Chloride-Thiourea-L-Tartaric acid
- Theoretical Prediction on the Structural, Electronic, Mechanical, and Thermodynamic Properties of TaSi2 with a C40 Structure Under Pressure
Articles in the same Issue
- Frontmatter
- General
- Analytical Solutions and Integrable Structure of the Time-Dependent Harmonic Oscillator With Friction
- Atomic, Molecular & Chemical Physics
- Wetting Behaviour of Gold Electrode and Molten Alkali Chlorides
- Probing Resonances and Pseudospin Symmetry of the Eckart Potential by the Complex Scaling Method within the Relativistic Framework
- Dynamical Systems & Nonlinear Phenomena
- First integrals and exact solutions of some compartmental disease models
- Gravitation & Cosmology
- Motion of a Test Particle According to the Scalar Ether Theory of Gravitation and Application to its Celestial Mechanics
- Hydrodynamics
- Analysis for Various Effects of Relaxation Time and Wall Properties on Compressible Maxwellian Peristaltic Slip Flow
- Acoustic Instability with Dynamic Charging in Quantum Plasmas
- Solid State Physics & Materials Science
- Preparation, Structural, Spectroscopic, Thermal, Linear and Nonlinear Optical Characteristics of Semi-Organic Material: Samarium Chloride-Thiourea-L-Tartaric acid
- Theoretical Prediction on the Structural, Electronic, Mechanical, and Thermodynamic Properties of TaSi2 with a C40 Structure Under Pressure