Materials Transactions, Vol. 51, No. 3 (2010) pp. 574 to 577 #2010 The Japan Institute of Metals RAPID PUBLICATION First-Principles Calculations of the Specific Heats of Cubic Carbides and Nitrides Satoshi Iikubo1 , Hiroshi Ohtani2 and Mitsuhiro Hasebe2 1 2 Graduate School of Life Science and Systems Engineering, Kyushu Institute of Technology, Kitakyushu 808-0196, Japan Department of Materials Science and Engineering, Kyushu Institute of Technology, Kitakyushu 804-8550, Japan Calculated specific heats of several carbides and nitrides with a B1 structure have been compared with those of experimental data up to 3000 K. The specific heats at constant pressure are calculated with the quasiharmonic approximation, using phonon dispersions calculated from direct method, and pseudopotential plane-wave method. The calculated results of HfC, TaC, TiC, ZrC, and ZrN are in excellent agreement with the experimental data up to 3000 K. For NbC, TiN, and VC, the calculated results are also in excellent agreement with the experimental data up to 2000 K, while they show an excessive rise over the experimental data above 2000 K. The deviation of the calculation from the experiment at high temperature is caused by instability of the B1 structure or anharmonic effect. [doi:10.2320/matertrans.MBW200913] (Received October 14, 2009; Accepted November 19, 2009; Published January 7, 2010) Keywords: specific heat, first-principles calculations, quasiharmonic approximation, carbide and nitride with B1 structure 1. Introduction The CALPHAD (CALculation of PHAse Diagrams) approach has been successfully used in a wide area of applications, from alloys to surfaces, interfaces, and clusters.1) In the current approach, the thermodynamic parameters necessary for calculating phase diagrams are estimated from the experimental data such as phase boundaries, activities and so on. It is difficult to estimate these parameters for metastable phases for which the physical and chemical properties are not experimentally accessible. One effective way to obtain information on the metastable phase is an electronic structure calculation based on the first-principles method.2) The application of the first-principles calculations to the study of thermodynamic properties at finite temperature remains challenging, because this calculation provides the thermodynamic properties at the ground state, i.e., zero temperature. Therefore, a method is required to calculate the specific heats for extending the ground state structure properties to those in a finite temperature region. The specific heat, which is closely related to the free energy, has been examined theoretically, both semi-empirically and with the first-principles. To calculate the specific heat at constant pressure, the anharmonic effect of the lattice vibration must be considered. The isopiestic specific heats at finite temperatures below 300 K have been discussed on the basis of the Debye and Grüneisen models to derive their thermal expansion.3) Using this model, Lu et al. show the thermodynamic properties of carbides and nitrides4,5) at temperatures up to 3000 K. An alternative approach for determining the lattice contribution to the free energy is the quasiharmonic approximation. Calculation based on this method suggests that the quasiharmonic approximation provides a reasonable description of the thermodynamic properties of simple metals,6,7) oxides,8–10) and nitrides11,12) below their melting points. Comparing these two methods, the former uses a fitting parameter to reproduce the experimental results in the extended temperature region, while the latter does not need any experimental data. Thus, the quasiharmonic approach matches our needs, and we attempted to calculate the specific heat, thermal expansion, and bulk modulus for HfC, NbC, TaC, TiC, TiN, VC, ZrC, and ZrN with the B1 structure, employing the direct method for obtaining phonon spectra. The essential results are recorded in this paper. 2. Theory and Calculation Procedure Then, the specific heat at constant volume is described by, @E ; ð1Þ Cv ¼ @T v where E denotes the internal energy. The experimental specific heat of a crystalline phase should be constructed from the electronic and lattice contributions. The internal energy due to the electron excitations Eel and phonon excitations Eph can be expressed as Z1 Z F el el D ðÞ f ð; TÞd Del ðÞd; ð2Þ E ðTÞ ¼ 1 1 Z1 Eph ðTÞ ¼ Dph ð!Þgð!; TÞh !d!; ð3Þ 0 where , F , Del ðÞ, f ð; TÞ, !, Dph ð!Þ, gð!; TÞ are electron energy, the Fermi energy, electronic density of state, Fermi distribution function, phonon frequency, the phonon density of states, and Bose distribution function, respectively. Then, the specific heat at constant pressure Cpph is given by, Cpph ¼ CVph þ v ðTÞ2 B0 ðTÞVðTÞT; ð4Þ where B0 ðTÞ is the bulk modulus and VðTÞ is the volume at temperature T. The coefficient of volume thermal expansion v ðTÞ is expressed as 1 @V : ð5Þ v ¼ V @T p To estimate the anharmonic lattice contribution included in Cpph , we used the quasiharmonic approximation. The vibrational free energy of the lattice ions F ph ðV; TÞ is XX h !j ðq; VÞ ln 2 sinh ; F ph ðV; TÞ ¼ EðVÞ þ kB T 2kB T q j ð6Þ First-Principles Calculations of the Specific Heats of Cubic Carbides and Nitrides where EðVÞ is the energy of the static lattice at a given volume V, and !j ðq; VÞ is the frequency of the jth phonon band at point q in the Brillouin zone. A quasiharmonic approximation, assuming that phonon frequencies depend only on the cell parameters, was employed here. This enabled us to take thermal expansion into account. To relate the vibrational free energy F ph ðV; TÞ and thermal properties, the third-order Birch–Murnaghan equation of state13) is used. #3 ( " 2 9V0 B0 V0 3 EðVÞ ¼ E0 þ 1 B00 16 V " 2 #2 " 2 #) V0 3 V0 3 þ 1 64 ; ð7Þ V V 575 (a) TaC (b) NbC -200 Free energy, F ph / (kJ/mol/F.U.) -100 -300 -400 3. Results and Discussion We first show the Helmholtz free energy calculated using the quasiharmonic approximation, and note a problem in estimating the anharmonic effect at higher temperature. Figure 1 shows a plot of the vibrational free energy F ph for TaC and NbC against a lattice volume at temperatures from 0 to 3000 K. The lattice volumes with minimum free energy were determined from fitting curves of the third-order Birch– Murnaghan equation of state at each temperature, given by the solid lines in Fig. 1. As shown in this figure, the lattice volume for both the compounds increases with increasing temperature. For TaC, the lattice volume of minimum free energy is 21.72 Å3 at 0 K, and substantially increase to 23.66 Å3 at 3000 K, whereas, for NbC, the figure shows that the lattice volume of the minimum free energy is 22.2 Å3 at 0 K and leaves the range of the figure above 2700 K, as shown in Fig. 1(b). It is difficult to estimate the minimum free energy at high temperature for NbC, because the system shows the anomalous behavior of the phonon dispersion. Figure 2 shows the phonon spectrum for NbC at (a) V0 ¼ 22:03 Å3 and (b) V0 ¼ 25:14 Å3 . Three acoustic and optical branches, separated by a frequency gap, are shown 20 21 22 23 24 21 3 22 23 24 3 Volume, V/Å Volume, V/Å Fig. 1 Temperature dependence of the free energy F on crystal volume is denoted by crosses for (a) TaC and (b) NbC at temperatures 0 < T < 3000 K at 100 K intervals. The dashed lines are results of fittings using the third-order Birch–Murnaghan equation of state. The solid line connects the F minima for different temperatures. 25 (b) NbC, V0 = 25.14 Å3 (a) NbC, V0 = 22.03 Å3 20 Frequency, ω/THz where V0 is the volume for the minimum energy, and B00 ¼ @B0 =@P is the pressure derivative of the bulk modulus. First-principles calculations were carried out using the Vienna Ab Initio Simulation Package (VASP),14) which is a calculation code based on the density functional theory for systems with periodic boundary conditions. To obtain the thermodynamic data at 0 K, the local density approximation with non spin polarization was applied and 520 eV was chosen for the plane-wave cutoff. The PHONON code,15) which calculates the phonon properties of a crystal by a direct method, was used to calculate the vibrational properties. In this method, the crystal supercell is first optimized by a full quantum mechanical electronic structure calculation. In such calculations, one can make a series of small displacements of one atom at a time and calculate the Hellmann–Feynman forces exerted on all atoms. The force constants are then derived from the ab initio Hellmann–Feynman forces. The first-principles and phonon calculations were carried out with 64 atoms for a perfect crystal of (TMX)4 (TM = transition metal, X ¼ C, N), which corresponds to a 2 2 2 unit cell of a fluorite structure. 15 10 5 0 W L Γ X W K W L Γ X W K Fig. 2 The phonon spectrum for NbC calculated at (a) V0 ¼ 22:03 Å3 and (b) V0 ¼ 25:14 Å3 along high symmetry directions in the Brillouin zone. in the figures. A softening of the acoustic modes near the L point can be observed in Fig. 2(b), consistent with the previous report.16) By introducing volume expansion to the system, the frequency of L point becomes an imaginary number, which is incompatible with the dynamic stability of NbC with the B1 structure. The same behavior is observed for VC, and TiN above 2500 K. Plausible causes of this problem might be the anharmonic effect of the crystal lattice, or a possible high-temperature structural transition, i.e., the order-disorder transition. For example, Isaev et al. reported structural instability of the B1 structure for a particular valence electron number, e.g., NbN.16) The calculated specific heats presented in Figs. 3(a)–3(h) are compared with the available experimental data,17,18) as shown by the circles. We note an excellent agreement between the theoretical and experimental results, especially for HfC, TaC, TiC, ZrC, and ZrN in a wide range of temperatures up to 3000 K. It can also be seen that below 1000 K, the difference between Cvph and Cpph is very small. At high temperature, Cvph approaches the classical constant value, while Cpph increases monotonously with the temper- ph Cp ph Cp + experiment 0 1000 2000 Temperature, T/K 80 (c) TiC 60 ph Cv 40 ph Cp ph Cp 20 0 Specific Heat, C/(J/mol/K) 0 1000 2000 Temperature, T/K (e) ZrN 60 ph Cv 40 ph Cp ph 20 el Cp + Cv experiment 0 1000 2000 Temperature, T/K (g) TiN 60 ph Cv ph Cp ph 20 el Cp + Cv experiment 0 0 1000 2000 Temperature, T/K 0 3000 1000 2000 Temperature, T/K 3000 80 (d) ZrC 60 ph Cv 40 ph Cp ph Cp 20 el + Cv experiment 0 0 1000 2000 Temperature, T/K 3000 80 (f) NbC TiC ZrN 80 300 HfC TaC 250 TaC 60 HfC TiC 200 ZrC ZrC 40 150 ZrN VC NbC NbC 100 20 TiN TiN 50 60 ph VC Cv 40 ph 0 Cp ph Cp 20 0 0 el Cv + experiment 2500 Tempelature, T/K 0 2500 Temperature, T/K 0 3000 80 40 0 3000 80 0 Specific Heat, C/(J/mol/K) el + Cv experiment Cv ph Cp ph el Cp + Cv experiment 20 3000 Specific Heat, C/(J/mol/K) Specific Heat, C/(J/mol/K) 0 Specific Heat, C/(J/mol/K) 20 el Cv ph 40 (b) (a) Bulk modulus, B0 /GPa ph Cv 40 350 (b) TaC 60 -1 60 80 -6 (a) HfC Linear thermal expansion coefficient, α l x10 /K 80 Specific Heat, C/(J/mol/K) S. Iikubo, H. Ohtani and M. Hasebe 0 Specific Heat, C/(J/mol/K) Specific Heat, C/(J/mol/K) 576 1000 2000 Temperature, T/K 3000 Fig. 4 Temperature dependence of (a) linear thermal expansion coefficient and (b) bulk modulus B0 ðTÞ for HfC, TaC, TiC, ZrC, ZrN, NbC, TiN, and VC, respectively. The curves are the calculated results and the symbols for represent experimental data from Ref. 19). 80 (h) VC 60 ph Cv 40 ph Cp ph 20 el Cp + Cv experiment 0 0 1000 2000 Temperature, T/K 3000 Fig. 3 Calculated temperature dependence of heat capacity of (a) HfC, (b) TaC, (c) TiC, (d) ZrC, (e) ZrN, (f) NbC, (g) TiN, and (h) VC at constant pressure (Cp ) and constant volume (Cv ). The experimental data for Cp (Ref. 13) and 14)) are denoted by circles. ature. The difference between the solid and dashed lines in Fig. 3 denotes the electronic contribution to the specific heats. It can readily be understood that the contribution is quite small, but not negligible. On the other hand, we see a marked difference in the high temperature region for NbC, TiN, and VC, which show the soft mode in the phonon calculation for a large volume. The calculated specific heats are much larger than the experimental ones above approximately 2000 K. This large discrepancy in high temperature is attributed to the anharmonic contribution or high-temperature structural instability. Figures 4(a) and (b) show the temperature dependence of the linear thermal expansion coefficient l and bulk modulus B0 obtained from the F–V curve fitting. In the regular system, as the temperature increase, the l and V are expected to increase, while B0 decreases. Available experimental results for linear thermal expansion coefficient l 19) are also indicated in this figure. Our calculations are in good agreement with the experimental data of l for HfC, TaC, TiC, ZrC, and ZrN. Thermal volume expansion can be indirectly reproduced by the quasiharmonic approximation. For TaC, we note a finite deviation between the calculation and experiment for l , while the specific heat shows good agreement with the experiment. A large deviation in l values for NbC, TiN, and VC is observed at high temperature. The difference between the calculated specific heats and exper- imental ones are clearly attributable to the over estimation of l and V, because the anharmonic contribution to isopiestic specific heat is described by a multiplication of these properties. These compounds also show anomalous softening of B0 with increasing temperature. The deviation in the l values and anomalous softening of B0 are obviously related to the anharmonic effect or high-temperature structural instability. 4. Conclusions The isopiestic specific heats for HfC, NbC, TaC, TiC, TiN, VC, ZrC, and ZrN with the B1 structure were calculated by taking the electron and lattice contributions into account at constant volume as well as an additional anharmonic component of the lattice vibration. The anharmonic part was described by the quasiharmonic approximation. The comparison between the experimental data and calculated results showed an excellent agreement even at high temperature. The present results show a possible development for the CALPHAD method that includes metastable structures that are not experimentally accessible. Acknowledgment The authors are grateful to Dr. S. Minamoto at CTC Technology Co. and Prof. I. Tanaka at Kyoto University for useful discussions. This work was partly supported by the Grant-in-Aid for Young Scientists (B) (21740239) from the Japanese Ministry of Education, Culture, Sports, Science and Technology. One author (H. O.) also acknowledges financial support by the Grant-in-Aid for Scientific Research on Priority Area, ‘‘Atomic Scale Modification’’ (Area No. 424) from the Japanese Ministry of Education, Culture, Sports, Science and Technology. First-Principles Calculations of the Specific Heats of Cubic Carbides and Nitrides REFERENCES 1) N. Saunders and A. P. Miodownik: CALPHAD (Calculation of Phase Diagrams): A Comprehensive Guide, Pergamon Materials Series, (Elsevier Science Ltd., Oxford, 1998). 2) For example, H. Ohtani, Y. Chen and M. Hasebe: Mater. Trans. 45 (2004) 1489–1498. 3) V. L. Moruzzi, J. F. Janak and K. Schwarz: Phys. Rev. B 37 (1988) 790–799. 4) X.-G. Lu, M. Selleby and B. Sundman: Acta Mater. 53 (2005) 2259– 2272. 5) X.-G. Lu, M. Selleby and B. Sundman: Acta Mater. 55 (2007) 1215– 1226. 6) J. Xie, S. Gironcoli, S. Baroni and M. Scheffler: Phys. Rev. B 59 (1999) 965–969. 7) A. Quong and A. Y. Liu: Phys. Rev. B 56 (1997) 7767–7770. 8) A. Kuwabara, T. Tohei, T. Yamamoto and I. Tanaka: Phys. Rev. B 71 (2005) 064301. 9) A. Seko, F. Oba, A. Kuwabara and I. Tanaka: Phys. Rev. B 72 (2005) 024107. 577 10) S. Minamoto, M. Kato, K. Konashi and Y. Kawazoe: J. Nucl. Mater. 385 (2009) 18–20. 11) T. Tohei, A. Kuwabara, F. Oba and I. Tanaka: Phys. Rev. B 73 (2006) 064304. 12) A. Kuwabara, K. Matsunaga and I. Tanaka: Phys. Rev. B 78 (2008) 064104. 13) J. P. Poirier: Introduction to the Physics of the Earth’s Interior, 2nd ed., (Cambridge University Press, New York, 2000). 14) G. Kresse and J. Furthmuller: Phys. Rev. B 54 (1996) 11169–11186. 15) K. Parlinski, Z. Q. Li and Y. Kawazoe: Phys. Rev. Lett. 78 (1997) 4063–4066. 16) E. I. Isaev, S. I. Simak, I. A. Abrikosov, R. Ahuja, Yu. Kh. Vekilov, M. I. Katsnelson, A. I. Lichtenstein and B. Johansson: J. Appl. Phys. 101 (2007) 123519. 17) JANAF thermodynamic tables, 3rd ed. J. Phys. Chem. Ref. Data 1985:14 (suppl. 1). 18) E. K. Storms: The refractory carbides, New York (NY): (Academic Press; 1967). 19) Y. S. Touloukian, R. K. Kirky, R. E. Taylor and T. Y. R. Lee, (ed.): Thermal properties of matter, TPRC data books, vol. 13. New York (NY), (Plenum Press; 1975).