Structural and anisotropic elastic properties of hexagonal MP (M = Ti ...

1 downloads 0 Views 2MB Size Report
PHiLosoPHiCaL MagaZine. 3657. Here, Etot refers to the total energy of the MP phase. Eiso(M) and Eiso(P) are the total energies of the isolated M and P atoms; ...
Philosophical Magazine, 2016 VOL. 96, NO. 35, 3654–3670 http://dx.doi.org/10.1080/14786435.2016.1234081

Structural and anisotropic elastic properties of hexagonal MP (M = Ti, Zr, Hf) monophosphides determined by first-principles calculations Runyue Li and Yonghua Duan School of Material Science and Engineering, Kunming University of Science and Technology, Kunming, China

ABSTRACT

First-principles calculations were performed to investigate the structural properties, phase stabilities, elastic properties and thermal conductivities of MP (M  =  Ti, Zr, Hf ) monophosphides. These monophosphides are thermodynamically and mechanically stable. Values for the bulk modulus B, shear modulus G, Young’s modulus E and Poisson’s ratio ν were calculated by Voigt–Reuss–Hill approximation. The mechanical anisotropy was discussed via several anisotropy indices and three-dimensional (3D) surface constructions. The order of elastic anisotropy is ZrP > HfP > TiP. The minimum thermal conductivities of these monophosphides were investigated using Clarke’s model and Cahill’s model. The results revealed that these monophosphides are suitable for use as thermal insulating materials and that their minimum thermal conductivities are anisotropic.

ARTICLE HISTORY

Received 31 March 2016 Accepted 5 September 2016 KEYWORDS

First-principles calculations; monophosphides; phase stability; anisotropic elasticity

1. Introduction Compounds composed of transition metals and main group elements are of interest in a variety of scientific areas including surface science [1], catalysis [2], astrophysics [3,4] and organometallic chemistry [5,6]. A large number of transition metal borides, carbides, nitrides and silicides have been extensively studied due to their excellent mechanical, chemical and electrical properties [7–9]. Some phosphides have remarkable physical and chemical properties; however, only a few reports of transition metal phosphides have been presented in the literature to date [10–12]. Among the transition metal phosphides, the metallic MP (M = Ti, Zr, Hf) monophosphides feature a high hardness [13–15]. TiP has a hexagonal structure [16], whereas ZrP and HfP possess cubic and hexagonal structures [17], respectively. The hexagonal Ti, Zr and Hf phosphides, termed β-(Ti, Zr and Hf)P, have a superstructure of NiAs that can be thought of as interpenetrating slabs of NaCl-like (ABC) and NiAs-like (BCB) sandwiches [18]. The cubic ZrP and HfP, termed α-ZrP and α-HfP, possess NaCl-type lattices [19]. α-ZrP has been successfully synthesised via a sodium co-reduction of ZrCl4 and PCl3 at 350 °C for 6 h [20]. In addition, an α-ZrP single crystal was obtained by chemical vapour deposition from CONTACT  Yonghua Duan 

[email protected]

© 2016 Informa UK Limited, trading as Taylor & Francis Group

Philosophical Magazine 

 3655

a system of ZrC14, PCl3, H2 and Ar on a quartz substrate at 1050–1250 °C [21]. Whiskers of β-HfP, α-ZrP and β-TiP have been prepared from HfC14/ZrCl4/TiC14 + PC13 + H2 + Ar gas mixtures at 1050–1250 °C using the mixed-metal impurity-activated chemical vapour deposition process [22]. β-ZrP could transform into α-ZrP at 1425 °C, and the relatively high temperature stabilities of the hexagonal phosphides follow the order TiP  0. As shown in Table 2, the elastic constants Cij of TiP, ZrP and HfP meet the mechanical stability criteria, indicating that these MP monophosphides are mechanically stable. The elastic constant C11 refers to the x directional (a-axis) resistance to linear compression, while the elastic constant C33 means the resistance to linear compression along the z direction (c-axis) [35]. Obviously, in Table 2, TiP possesses the largest C11, suggesting that TiP is the most incompressible along the a-axis. ZrP has the smallest C11, which corresponds to a large compressibility of ZrP under a-axis uniaxial stress. Additionally, it is noted that the order of C33 is similar to that of C11. As a result, the order of incompressibility along the c-axis is also ZrP 0.26) is ductile. Otherwise, it is brittle [41,42]. As shown in Table 3, it is obvious that TiP and HfP are brittle due to their small values of B/G and ν, less than 1.75 and 0.26, respectively. ZrP has critical values of exactly B/G = 1.75 and ν = 0.26. However, it can be further explored using the classical criteria of the Cauchy pressure (C12−C44). A crystal with a negative Cauchy pressure is brittle; otherwise, it is ductile [43]. The negative values of the Cauchy pressure for TiP (−57.9 GPa), ZrP (−25.3 GPa) and HfP (−58.8 GPa) confirm their brittle nature. HfP is the most brittle phase, whereas ZrP is the least brittle phase. The hardness H is a basic parameter reflecting the elastic and plastic properties of a solid. The hardness of polycrystalline materials can be evaluated using a theoretical model [44]:

( )0.583 H = 2 k2 G − 3,

(11)

where k refers to Pugh’s modulus ratio (k = G/B). The hardness calculated from the theoretical model may be not very accurate because of the macroscopic concepts of bulk and shear moduli. However, it can provide a guideline to reveal basal factors controlling the hardness of solids. The calculated results of hardness for the considered monophosphides are also listed in Table 3. It can be seen that the calculated and experimental values of the hardness of TiP are 14.97 and 13.31 GPa [10], respectively, indicating that the calculated value agrees with the experimental one. Figure 3 plots the values of the elastic moduli (B, G and E) and hardness H for TiP, ZrP and HfP. For better comparison with the elastic moduli, the values of H are multiplied by a factor of 10. In Equation (11), the hardness is related to both the bulk and shear moduli. However, from Figure 3, the shear modulus is more pertinent to hardness than the bulk modulus due to the different natures of bulk and shear moduli. TiP has the highest hardness compared to the other monophosphides, which is in good agreement with the result predicted from shear modulus. There are no available experimental and theoretical data of the elastic moduli and the hardness for TiP, ZrP and HfP. However, a useful guide for future work can be provided by our calculations. 3.3.  Anisotropy of elastic moduli The elastic anisotropy is of importance to the possibility of including micro-cracks in crystals. There are several methodologies used to indicate the elastic anisotropy of hexagonal crystals. First, the elastic anisotropy of MP monophosphides can be characterised by the values of the bulk modulus along the a-axis (Ba) and c-axis (Bc), and the expressions are as follows [45],

Philosophical Magazine 

 3661

Figure 3. (colour online) Calculated bulk, shear and Young’s moduli and hardness of TiP, ZrP and HfP. The values of hardness are multiplied by the factor of 10 for a better comparison.

Ba = a

Λ dP = , da 2+𝛼

(12)

B dP = a, dc 𝛼

(13)

Bc = c

Λ = 2(C11 + C12 ) + 4C13 𝛼 + C33 𝛼 2 ,

𝛼=

C11 + C12 − 2C13 . C33 − C13

(14)

(15)

Second, the elastic anisotropy can be described by the universal anisotropy index AU and by the anisotropy in compression and shear Acomp and Ashear, respectively. The elastic anisotropy indexes AU, Acomp and Ashear for a crystal with any symmetry can be expressed as follows [46,47]:

AU = 5

Acomp =

GV BV + − 6 ≥ 0, GR BR

B V − BR , BV + BR

Ashear =

G V − GR GV + GR

(16)

(17)

where BV (GV) and BR (GR) are the bulk moduli (shear moduli) in the Voigt and Reuss approximations, respectively. A crystal with AU = 0 is isotropic. The large deviation of AU from zero suggests a high mechanical anisotropy. The values of Acomp and Ashear ranging from zero to 1 represent isotropy and the maximum elastic anisotropy [48].

3662 

  R. LI AND Y. DUAN

Third, for hexagonal crystals, the shear anisotropic factors A1, A2 and A3 are calculated according to the following expressions [49],

A1 =

4C44 C11 + C33 − 2C13

for the {1 0 0} plane

(18)

A2 =

4C55 C22 + C33 − 2C23

for the {0 1 0} plane

(19)

A3 =

4C66 C11 + C22 − 2C12

for the {0 0 1} plane

(20)

A crystal with A1, A2 and A3 values equal to 1.0 is isotropic. Otherwise, it is anisotropic. Finally, by solving the Christofel equation for a hexagonal lattice, the anisotropy of the compression wave (P) is obtained by [50]

ΔP =

C33 C11

(21)

The anisotropies of the shear wave polarised perpendicular to the basal plane (S1) and that polarised in the basal plane (S2) are calculated by

ΔS1 = ΔS2 =

C11 + C33 − 2C13 4C44 2C44

.

(22)

C11 − C12

For S2 and P waves, the extreme values occur along the c-axis; for S1, the extreme value occurs at an angle of 45° from the c-axis in the ac plane. A crystal with ΔP = ΔS1 = ΔS2 = 1 is isotropic. Several elastic anisotropy indexes are listed in Table 4. It can be observed that Bc is larger than Ba for these monophosphides. The values of Bc/Ba are 1.398, 1.450 and 1.425 for TiP, ZrP and HfP, respectively. This indicates that the compression along the c-axis is more difficult than that along the a-axis. This result agrees well with the comparison of C11 and C33. This can be explained by the presence of the ionic metal–phosphorus bonds in the [0 0 1] direction in these monophosphides. In Table 4, it is clear that the order of the value of AU is TiP  ZrP > HfP. These MP monophosphides are ideally qualified for use as thermal insulating materials because of their relatively low thermal conductivity.

Philosophical Magazine 

 3667

Because the total thermal conductivity is treated as the summation of three acoustic branches (one is the longitudinal mode and the others are transverse modes), Cahill’s model should fit better than Clarke’s model for a discussion of the anisotropy in thermal conductivity [62]. According to the symmetry of hexagonal crystal, the results listed in Table 6 are confined to two low index crystallographic directions, ie the [1 0 0] and [0 0 1] directions. The values of vl, vt1 and vt2 in the [1 0 0] and [0 0 1] directions in Table 5 are significantly different. This results in the difference of the calculated thermal conductivities (λmin[1 0 0] and λmin[0 0 1]) in Table 6. λmin[0 0 1] always being larger than λmin[1 0 0] in these MP monophosphides suggests that the dependences of thermal conductivities in the [0 0 1] direction are more prominent than those in the [1 0 0] direction. In hexagonal monophosphides, M atoms are aligned mainly along the [0 0 1] direction, whereas P atomic layers are parallel to the (1 0 0) plane. The phonon collision probability along the [1 0 0] direction is smaller than that along the [0 0 1] direction. Thus, the phonon mean free path along the [1 0 0] direction is larger than that along the [0 0 1] direction. As a result, the thermal conductivity in the [0 0 1] direction is larger than that in the [1 0 0] direction.

4. Conclusions First-principles methods have been utilised to systematically investigate the phase stability, elastic moduli, hardness, elastic anisotropy properties, Debye temperatures and thermal conductivity of MP (M = Ti, Zr, Hf) monophosphides. The obtained formation enthalpies indicate that HfP is the most stable phase. The calculated elastic constants reveal that these monophosphides are mechanically stable. The bulk modulus, shear modulus, Young’s modulus, Poisson’s ratio and hardness were calculated and discussed. The values of C33 for MP are very large among the elastic constants, which correspond to an incompressibility of MP under z direction (c-axis) uniaxial stress. According to several anisotropy indexes and 3D surface constructions, the order of the elastic anisotropy of MP monophosphides is ZrP > HfP > TiP. Furthermore, the order of the Debye temperature and the minimum thermal conductivity for these monophosphides is TiP > ZrP > HfP, and these monophosphides are suitable for use as thermal insulating materials. The thermal conductivity and the sound velocity are anisotropic in these crystal structures based on calculated sound velocities and thermal conductivities. We hope our results can provide guidance for further theoretical and experimental studies on these monophosphides.

Disclosure statement No potential conflict of interest was reported by the authors.

Funding This work was supported by the Reserve Talents Project of Yunnan Province [grant number 2015HB019], the National Natural Science Foundation of China [grant number 51201079] and the Scientific Research Fund of the Department of Education of Yunnan Province [grant number 2015Z038].

3668 

  R. LI AND Y. DUAN

References   [1] M. Wojciechowska, J. Haber, S. Lomnicki, and J. Stoch, Structure and catalytic activity of double oxide system: Cu–Cr–O supported on MgF2, J. Mol. Catal. A: Chem. 141 (1999), pp. 155–170. http://dx.doi.org/10.1016/S1381-1169(98)00259-3   [2] C.N.R. Rao, Transition metal oxides, Anu. Rev. Phys. Chem. 40 (1989), pp. 291–326. http:// dx.doi.org/10.1146/annurev.pc.40.100189.001451  [3]  W. Welter, Stellar and other high-temperature molecules, Science 155 (1967), pp. 155–164. http:// dx.doi.org/10.1126/science.155.3759.155   [4] N.M. White and R.F. Wing, Photoelectric two-dimensional spectral classification of M supergiants, Astrophys. J. 222 (1978), pp. 209–219.  [5]  A. Veillard, Ab initio calculations of transition-metal organometallics: Structure and molecular properties, Chem. Rev. 91 (1991), pp. 743–766. http://dx.doi.org/10.1021/cr00005a006   [6] N. Koga and K. Morokuma, Ab initio molecular orbital studies of catalytic elementary reactions and catalytic cycles of transition-metal complexes, Chem. Rev. 91 (1991), pp. 823–842. http:// dx.doi.org/10.1021/cr00005a010   [7] A.J. Merer, Spectroscopy of the diatomic 3d transition metal oxides, Annu. Rev. Phys. Chem. 40 (1989), pp. 407–438. http://dx.doi.org/10.1146/annurev.pc.40.100189.002203   [8] M. Hargittai, Molecular structure of metal halides, Chem. Rev. 100 (2000), pp. 2233–2302. http://dx.doi.org/10.1021/cr970115u   [9] B. Huang, Y.H. Duan, W.C. Hu, Y. Sun, and S. Chen, Structural, anisotropic elastic and thermal properties of MB (M = Ti, Zr and Hf) monoborides, Ceram. Int. 41 (2015), pp. 6831–6843. http:// dx.doi.org/10.1016/j.ceramint.2015.01.132 [10] R.L. Riplay, The preparation and properties of some transition phosphides, J. Less Common Met. 4 (1962), pp. 496–503. http://dx.doi.org/10.1016/0022-5088(62)90037-1 [11] M. Knausenberger, G. Brauer, and I.A. Gingerich, Preparation and phase studies of titanium phosphides, J. Less Common Met. 8 (1965), pp. 136–148. http://dx.doi.org/10.1016/00225088(65)90105-0 [12] S. Motojima, T. Wakamatsu, Y. Takahashi, and K. Sugiyama, Crystal growth and some properties of titanium monophosphide, J. Electrochem. Soc. 123 (1976), pp. 290–295. http://dx.doi. org/10.1149/1.2132805 [13] E.F. Strotzer, W. Biltz, and K. Meisel, Beiträge zur systematischen Verwandtschaftslehre. 84. Zirkoniumphosphide [Contributions to the systematic affinity. 84. Zirkoniumphosphide], Z. Anorg. Chem. 239 (1938), pp. 216–224. http://dx.doi.org/10.1002/zaac.19382390211 [14] W. Blitz, A. Rink, and F. Wiechman, Compounds of Ti with P; Contribution to systematic affinity theory, Z. Anorg. Chem. 238 (1938), pp. 395–405. [15]  K.A. Gingerich, Stability and vaporization behaviour of group IV–VI transition metal monophosphides, Nature 200 (1963), pp. 877–877. http://dx.doi.org/10.1038/200877a0 [16] N. Schönberg, An X-ray investigation of transition metal phosphides, Acta Chem. Scand. 8 (1954), pp. 226–239. [17] K.S. Irani and K.A. Gingerich, Structural transformation of zirconium phosphide, J. Phys. Chem. Solids 24 (1963), pp. 1153–1158. http://dx.doi.org/10.1016/0022-3697(63)90231-2 [18] P.-O. Snell, The crystal structure of TiP, Acta Chem. Scand. 27 (1967), pp. 1773–1776. [19] R.F. Jarvis Jr., R.M. Jacubinas, and R.B. Kaner, Self-propagating metathesis routes to metastable group 4 phosphides, Inorg. Chem. 39 (2000), pp. 3243–3246. http://dx.doi.org/10.1021/ ic000057m [20] L. Chen, M. Huang, Y. Gu, L. Shi, Z. Yang, and Y. Qian, Low-temperature synthesis of nanocrystalline ZrP via co-reduction of ZrCl4 and PCl3, Mater. Lett. 58 (2004), pp. 3337–3339. http://dx.doi.org/10.1016/j.matlet.2004.06.033 [21] S. Motojima, Y. Takahashi, and K. Sugiyama, Chemical vapour deposition of zirconium phosphide whiskers, J. Cryst. Growth 30 (1975), pp. 1–8. http://dx.doi.org/10.1016/0022-0248(75)90191-8 [22] M. Fujii, H. Iwanaga, and S. Motojima, CVD growth and morphology of transition-metal phosphides, J. Cryst. Growth 166 (1996), pp. 99–103. http://dx.doi.org/10.1016/0022-0248(95)00525-0

Philosophical Magazine 

 3669

[23] T.S. Lewkebandara, J.W. Proscia, and C.H. Winter, Precursor for the low-temperature deposition of titanium phosphide films, Chem. Mater. 7 (1995), pp. 1053–1054. http://dx.doi.org/10.1021/ cm00054a003 [24] C.S. Blackman, C.J. Carmalt, I.P. Parkin, S.A. O’Neill, K.C. Molloy, and L. Apostolico, Chemical vapour deposition of crystalline thin films of tantalum phosphide, Mater. Lett. 57 (2003), pp. 2634–2636. http://dx.doi.org/10.1016/S0167-577X(02)01341-1 [25] H.-J. Sue, K.T. Gam, N. Bestaoui, N. Spurr, and A. Clearfield, Epoxy nanocomposites based on the synthetic α-zirconium phosphate layer structure, Chem. Mater. 16 (2004), pp. 242–249. http://dx.doi.org/10.1021/cm030441s [26] M. Pica, A. Donnadio, and M. Casciola, Starch/zirconium phosphate composite films: Hydration, thermal stability, and mechanical properties, Starch-Stärke 64 (2012), pp. 237–245. http://dx.doi. org/10.1002/star.201100113 [27] M. Hussain, Y.-H. Ko, and Y.-H. Choa, Significant enhancement of mechanical and thermal properties of thermoplastic polyester elastomer by polymer blending and nanoinclusion, J. Nanomater. 2016 (2016), pp. 1–9. http://dx.doi.org/10.1155/2016/8515103 [28] M.D. Segall, P.J.D. Lindan, M.J. Probert, C.J. Pickard, P.J. Hasnip, S.J. Clark, and M.C. Payne, First-principles simulation: Ideas, illustrations and the CASTEP code, J. Phys.: Condens. Matter. 14 (2002), pp. 2717–2744. http://dx.doi.org/10.1088/0953-8984/14/11/301 [29] J.P. Perdew, K. Burke, and Y. Wang, Generalized gradient approximation for the exchangecorrelation hole of a many-electron system, Phys. Rev. B 54 (1996), pp. 16533–16539. http:// dx.doi.org/10.1103/PhysRevB.54.16533 [30] K. Bachmayer, H. Nowotny, and A. Kohl, Die struktur von TiAs [The structure of TiAs], Monatsh. Chem, 86 (1955), pp. 39–43. http://dx.doi.org/10.1007/BF00899271 [31] W. Jeitschko and H. Nowotny, Die struktur von HfP [The structure of HfP], Monatsh. Chem. 93 (1962), pp. 1107–1109. http://dx.doi.org/10.1007/BF00905910 [32] G. Hägg, Gesetzmässigkeiten im kristallbau bei hydriden, boriden, carbiden und nitriden der übergangselemente [Laws in crystal construction in hydrides, borides, carbides, nitrides of transition elements], Z. Phys. Chem. 12 (1931), pp. 33–56. [33] C.J. Qi, J. Feng, R.F. Zhou, Y.H. Jiang, and R. Zhou, First principles study on the stability and mechanical properties of MB (M = V, Nb and Ta) compounds, Chin. Phys. Lett. 30 (2013), pp. 117101-1–117101-5. http://dx.doi.org/10.1088/0256-307X/30/11/117101 [34] F. Birch, Finite strain isotherm and velocities for single-crystal and polycrystalline NaCl at high pressures and 300°K, J. Geophys. Res. 83 (1978), pp. 1257–1268. http://dx.doi.org/10.1029/ JB083iB03p01257 [35] X.P. Gao, Y.H. Jiang, R. Zhou, and J. Feng, Stability and elastic properties of Y-C binary compounds investigated by first principles calculations, J. Alloys Compd. 587 (2014), pp. 819–826. http:// dx.doi.org/10.1016/j.jallcom.2013.11.005 [36] H. Ozisik, E. Deligoz, K. Colakoglu, and G. Surucu, Structural and mechanical stability of rareearth diborides, Chin. Phys. B 4 (2013), pp. 046202-1–046202-8. http://dx.doi.org/10.1088/16741056/22/4/046202 [37] W. Voigt, Lehrbuch der Kristallphysik [Handbook on crystal physics], Taubner, Leipzig, 1928. [38] A. Reuss, Berechnung der fließgrenze von mischkristallen auf grund der plastizitätsbedingung für einkristalle [Calculation of the yield of mixed crystals on the basis of the plasticity condition for single crystals], Z. Angew. Math. Mech. 9 (1929), pp. 49–58. http://dx.doi.org/10.1002/ zamm.19290090104 [39] R. Hill, The elastic behaviour of a crystalline aggregate, Proc. Phys. Soc. A 65 (1952), pp. 349–354. http://dx.doi.org/10.1088/0370-1298/65/5/307 [40] N. Korozlu, K. Colakoglu, E. Deligoz, and S. Aydin, The elastic and mechanical properties of MB12 (M = Zr, Hf, Y, Lu) as a function of pressure, J. Alloys Compd. 546 (2013), pp. 157–164. http://dx.doi.org/10.1016/j.jallcom.2012.08.062 [41] S.F. Pugh, XCII, Relations between the elastic moduli and the plastic properties of polycrystalline pure metals, Philos. Mag. 45 (1954), pp. 823–843. http://dx.doi.org/10.1080/14786440808520496 [42] J.J. Lewandowski, W.H. Wang, and A.L. Greer, Intrinsic plasticity or brittleness of metallic glasses, Philos. Mag. Lett. 85 (2005), pp. 77–87. http://dx.doi.org/10.1080/09500830500080474

3670 

  R. LI AND Y. DUAN

[43] S. Chen, Y. Sun, Y.H. Duan, B. Huang, and M.J. Peng, Phase stability, structural and elastic properties of C15-type Laves transition-metal compounds MCo2 from first-principles calculations, J. Alloys Compd. 630 (2015), pp. 202–208. http://dx.doi.org/10.1016/j.jallcom.2015.01.038 [44] X.Q. Chen, H.Y. Niu, D.Z. Li, and Y.Y. Li, Modeling hardness of polycrystalline materials and bulk metallic glasses, Intermetallics 19 (2011), pp. 1275–1281. http://dx.doi.org/10.1016/ j.intermet.2011.03.026 [45] A.K.M.A. Islam, A.S. Sikder, and F.N. Islam, NbB2: A density functional study, Phys. Lett. A 350 (2006), pp. 288–292. http://dx.doi.org/10.1016/j.physleta.2005.09.085 [46] J. Feng, B. Xiao, R. Zhou, W. Pan, and D.R. Clarke, Anisotropic elastic and thermal properties of the double perovskite slab–rock salt layer Ln2SrAl2O7 (Ln = La, Nd, Sm, Eu, Gd or Dy) natural superlattice structure, Acta Mater. 60 (2012), pp. 3380–3392. http://dx.doi.org/10.1016/j. actamat.2012.03.004 [47] S.I. Ranganathan and M. Ostoja-Starzewski, Universal elastic anisotropy index, Phys. Rev. Lett. 101 (2008), pp. 055504. http://dx.doi.org/10.1103/PhysRevLett.101.055504 [48] H.B. Ozisik, K. Colakoglu, and E. Deligoz, First-principles study of structural and mechanical properties of AgB2 and AuB2 compounds under pressure, Comput. Mater. Sci. 51 (2012), pp. 83–90. http://dx.doi.org/10.1016/j.commatsci.2011.07.043 [49] P. Ravindran, L. Fast, P.A. Korzhavyi, and B. Johansson, Density functional theory for calculation of elastic properties of orthorhombic crystals: Application to TiSi2, J. Appl. Phys. 84 (1998), pp. 4891–4904. http://dx.doi.org/10.1063/1.368733 [50] G. Steinle-Neumann, L. Stixtude, and R.E. Cohen, First-principles elastic constants for the hcp transition metals Fe Co, and Re at high pressure, Phys. Rev. B 60 (1999), pp. 791–799. http:// dx.doi.org/10.1103/PhysRevB.60.791 [51] M. Born and K. Huang, Dynamical theory of crystal lattices, Clarendon, Oxford, 1954. [52] J.F. Nye, Physical Properties of Crystals, Oxford University Press, Oxford, 1985. [53] K. Brugger, Determination of third-order elastic coefficients in crystals, J. Appl. Phys. 36 (1965), pp. 768–773. http://dx.doi.org/10.1063/1.1714216 [54] D. Music, A. Houben, R. Dronskowski, and J.M. Schneider, Ab initio study of ductility in M2AlC (M = Ti, V, Cr), Phys. Rev. B 75 (2007), pp. 174102-1–174102-5. http://dx.doi.org/10.1103/ PhysRevB.75.174102 [55] J. Feng, B. Xiao, C.L. Wan, Z.X. Qu, Z.C. Huang, J.C. Chen, R. Zhou, and W. Pan, Electronic structure, mechanical properties and thermal conductivity of Ln2Zr2O7 (Ln = La, Pr, Nd, Sm, Eu and Gd) pyrochlore, Acta Mater. 59 (2011), pp. 1742–1760. http://dx.doi.org/10.1016/ j.actamat.2010.11.041 [56] Y. Shen, D.R. Clarke, and P.A. Fuierer, Anisotropic thermal conductivity of the aurivillus phase, bismuth titanate (Bi4Ti3O12): A natural nanostructured superlattice, Appl. Phys. Lett. 93 (2008), pp. 102907-1–102907-3. http://dx.doi.org/10.1063/1.2975163 [57] D.R. Clarke, Materials selection guidelines for low thermal conductivity thermal barrier coatings, Surf. Coat. Technol. 163–164 (2003), pp. 67–74. http://dx.doi.org/10.1016/S02578972(02)00593-5 [58] D.R. Clarke and C.G. Levi, Materials design for the next generation thermal barrier coatings, Annu. Rev. Mater. Res. 33 (2003), pp. 383–417. http://dx.doi.org/10.1146/annurev. matsci.33.011403.113718 [59] D.G. Cahill, S.K. Watson, and R.O. Pohl, Lower limit to the thermal conductivity of disordered crystals, Phys. Rev. B 46 (1992), pp. 6131–6140. http://dx.doi.org/10.1103/PhysRevB.46.6131 [60] Y.H. Duan, Y. Sun, and L. Lu, Thermodynamic properties and thermal conductivities of TiAl3-type intermetallics in Al–Pt–Ti system, Comput. Mater. Sci. 68 (2013), pp. 229–233. http://dx.doi. org/10.1016/j.commatsci.2012.11.012 [61] J. Callaway, Model for lattice thermal conductivity at low temperatures, Phys. Rev. 113 (1959), pp. 1046–1051. http://dx.doi.org/10.1103/PhysRev.113.1046 [62] J. Feng, B. Xiao, R. Zhou, and W. Pan, Anisotropy in elasticity and thermal conductivity of monazite-type REPO4 (RE = La, Ce, Nd, Sm, Eu and Gd) from first-principles calculations, Acta Mater. 61 (2013), pp. 7364–7383. http://dx.doi.org/10.1016/j.actamat.2013.08.043

Suggest Documents