Abstract
First-principles calculations were carried out to explore the structural stability, elastic moduli, ductile or brittle behaviour, anisotropy, dynamical stability, and thermodynamic properties of pure Al and CeT2Al20 (T = Ti, V, Cr, Nb, and Ta) intermetallics. The calculated formation enthalpy and phonon frequencies confirm that these intermetallics satisfy the conditions for structural stability. The elastic constants Cij, elastic moduli B, G, and E, and the hardness Hv indicate these intermetallics have higher hardness and the better resistance against deformation than pure Al. The values of Poisson’s ratio (v) and B/G indicate that CeT2Al20 intermetallics are all brittle materials. The anisotropic constants and acoustic velocities confirm that CeT2Al20 intermetallics are all anisotropic, but CeV2Al20, CeNb2Al20, and CeTa2Al20 are nearly isotropic. Importantly, the calculated thermodynamic parameters show that CeT2Al20 intermetallics exhibit better thermodynamic properties than pure Al at high temperature.
1 Introduction
Aluminium alloys have been playing important roles in many industrial fields because of their outstanding physical properties [1]. However, their poor creep resistance has hindered their wide applications at high temperature. In the past few decades, many researchers have found that the addition of appropriate transition-metal (TM) and rare earth (RE) elements to aluminium alloys can form L12-, D022-, D023-, or D019-ordered binary Al3TM and Al3RE intermetallics. These binary intermetallics possess some outstanding physical properties, such as low mass density, high melting point, good oxidation resistance, and so on, which can effectively improve the poor creep resistance of aluminium alloys [2], [3], [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15]. In order to further expand the applications of aluminium alloys, some A-T-Al ternary systems (where A = actinide/lanthanide/calcium; T = transition metal) were found that can form cubic AT2Al20 intermetallics with the CeCr2Al20-type structure [16], [17], [18], [19], [20], [21], [22], [23], [24], [25], [26], [27], [28], [29], [30], [31], [32]. Kangas and Thiede synthesised these new AT2Al20 intermetallics, namely CeT2Al20 (T = Ti, V, Cr, Nb and Ta), by using a step-wise arc-melting technique and by the self-flux method [31], [32]. In the case of these new CeT2Al20 intermetallics, the current interest is mainly focused on their synthesis and characterisation, but their mechanical properties, ductile or brittle behaviour, dynamic stability, and thermodynamic properties have not been completely clarified. It is necessary to systematically investigate their physical properties for further experimental studies and applications. In this work, we used first-principles calculations to comparatively investigate the structural stability, mechanical and dynamic stability, and thermodynamic properties of Al and five CeT2Al20 intermetallics.
2 Computational Methods
All first-principles calculations were based on the density functional theory (DFT) method in conjunction with projector augmented wave (PAW) potentials, within generalized gradient approximation (GGA) of PBE as implemented in the VASP software. The energy cut-off used was 600 eV for the plane wave basis for all compounds. The Monkhorst-Pack k mesh of 9 × 9 × 9 was selected for the Brillouin zone sampling. The phonon dispersions, density of states, and thermodynamic quantities were calculated by using the Phonopy code [33] together with the VASP code. The supercell method and linear response method were used for phonon calculations. The stress-strain method was used to calculate the elastic constants of CeT2Al20 intermetallics [34].

Crystal structure of CeTAl20 intermetallics. The violet balls are Ce atoms, the red balls are T atoms, and the blue balls are the Al atoms.
Lattice constant a0 (Å), bulk modulus B0 (GPa), and formation enthalpy ΔH (eV/atom) of CeT2Al20 (T = Ti, V, Cr, Nb, and Ta) intermetallics.
Compounds | a0 | B0 | ΔH | Reference |
---|---|---|---|---|
CeTi2Al20 | 14.927 | 83.263 | −0.3283 | This work |
14.710 | Exp. [31] | |||
CeV2Al20 | 14.856 | 78.361 | −0.3318 | This work |
14.552 | Exp. [31] | |||
CeCr2Al20 | 14.639 | 89.214 | −0.337 | This work |
14.484 | Exp. [32] | |||
CeNb2Al20 | 15.069 | 88.967 | −0.3359 | This work |
14.748 | Exp. [32] | |||
CeTa2Al20 | 15.085 | 89.425 | −0.3362 | This work |
14.748 | Exp. [32] |
3 Results and Discussion
3.1 Structural Stability
CeT2Al20 intermetallics can crystallise in face-centred cubic structure with the Fd-3m space group. In the prototypical CeT2Al20 compounds, Ce atoms occupy the 8a Wyckoff sites, the T atoms occupy the 16d Wyckoff sites, and the Al atoms occupy the 16c, 48f, and 96g Wyckoff sites. The unit cell of CeT2Al20 intermetallics is shown in Figure 1 [25]. In order to investigate the structural properties in the ground state, the total energy under different volumes was calculated and fitted to the Birch-Murnaghan equation of state (EoS). The obtained lattice constants and bulk moduli are listed in Table 1. In this table, the calculated values are slightly larger than the experimental ones [31], [32] but the error is less than 5 %, which indicates that our computational method is reliable. CeNb2Al20 has the largest bulk modulus B0 among the five intermetallics, which indicates that CeNb2Al20 has the highest resistance against compression and volume change. In order to obtain the structural stability under high pressure, Figure 2 displays the relationship between the values of V/V0 under different pressures. These curves are smooth and have no inflection points until 50 GPa (Fig. 2), indicating that these intermetallics show structural stability under high pressure. The curves of CeTa2Al20 and CeV2Al20 are located at the upper and lower part, respectively, which indicates that CeTa2Al20 has the highest resistance to compression and CeV2Al20 has the lowest, among the five intermetallic compounds.

Equation of state for CeT2Al20 intermetallics.

Phonon dispersion curves of Al; the circles correspond to experimental data.

Phonon dispersion curves and density of states for CeT2Al20 intermetallics.
Generally speaking, the structural stability of solid materials mainly depends on two factors: the thermodynamic factor, which depends on the formation enthalpy; and the dynamic factor, which depends on the phonon dispersion and density of states [35], [36]. The formation enthalpy of CeT2Al20 intermetallics was calculated using the following expression [37], [38], [39], [40]:
where ETotal represents the total energy of CeT2Al20; x, y, and z are the number of atoms in the unit cell; ECe, ET, and EAl represent the total energy per atom in the solid. The calculated formation enthalpies (
In order to explore their dynamic stability, the phonon dispersion and density of states were calculated, which are shown in Figures 3 and 4. Figure 3 shows the phonon dispersion curves of pure Al. The solid lines are the calculated values and the scattered points are the experimental ones obtained from inelastic neutron scattering measurements [41]. The calculated values are very close to experimental ones, indicating that the present method and the results are reasonable. Figure 4 shows the phonon dispersion curves and density of states for the five CeT2Al20 intermetallic compounds. No imaginary phonon frequency is present along the Brillouin zone path in Figure 4, which indicates that the five intermetallic compounds have dynamic stability. The masses of the five intermetallics follow the order: CeTi2Al20 < CeV2Al20 < CeCr2Al20 < CeNb2Al20 < CeTa2Al20, and the maximum values of the phonon frequency gradually decrease with increasing mass. In the phonon density of states (PHDOS) also, imaginary frequencies are absent, which also confirms that the five intermetallics have dynamic stability. In summary, the calculated formation enthalpies and phonon frequencies confirm that the five CeT2Al20 intermetallics are structurally stable.
Calculated elastic constants Cij (GPa) and compliance matrices Sij.
Model | C11 | C12 | C44 | S11 | S12 | S44 |
---|---|---|---|---|---|---|
CeTi2Al20 | 146.745 | 48.349 | 66.080 | 0.0081 | −0.0020 | 0.0143 |
CeV2Al20 | 159.992 | 34.119 | 64.609 | 0.0067 | −0.0012 | 0.0155 |
CeCr2Al20 | 139.913 | 58.948 | 52.438 | 0.0095 | −0.0028 | 0.0191 |
CeNb2Al20 | 168.828 | 43.675 | 67.406 | 0.0066 | −0.0014 | 0.0148 |
CeTa2Al20 | 172.426 | 43.992 | 70.893 | 0.0065 | −0.0013 | 0.0145 |
Al | 112.091 | 63.424 | 32.065 | 0.0151 | −0.0055 | 0.0312 |
Al [42] | 124 | 61.9 | 31.6 | – | – | – |
Al [43] | 109 | 64.5 | 32.4 | – | – | – |
Al [44] | 108 | 62 | 28 | – | – | – |
3.2 Elastic Properties
The elastic constants and elastic moduli are important parameters for engineering applications. Table 2 lists the calculated elastic constants (Cij) and compliance matrix constants (Sij) for CeT2Al20 intermetallics and pure Al [42], [43], [44]. The values of the elastic constants of Al are consistent with the experimental ones, indicating that the computational methods are reliable. For these cubic structures, there are three independent constants, namely C11, C12, and C44. For mechanical stability, the elastic constants must satisfy the following conditions: C11 > 0, C44 > 0,
After obtaining the elastic constants Cij and Sij, the bulk modulus B, the shear modulus G, Young’s modulus E, and the hardness Hv were calculated by using the VRH method and the following equations [46], [47]:
Calculated elastic moduli B, G, E, Hv (GPa), G/B, Poisson’s ratio v, and melting temperature (K).
Model | B | G | E | B/G | Hv | υ | Tm ± 300 |
---|---|---|---|---|---|---|---|
CeTi2Al20 | 81.148 | 60.818 | 145.984 | 1.334 | 12.138 | 0.200 | 1420.263 |
CeV2Al20 | 76.077 | 63.935 | 149.831 | 1.189 | 12.827 | 0.172 | 1498.553 |
CeCr2Al20 | 85.937 | 47.277 | 119.852 | 1.818 | 7.149 | 0.268 | 1379.884 |
CeNb2Al20 | 85.393 | 65.431 | 156.357 | 1.305 | 13.114 | 0.195 | 1550.773 |
CeTa2Al20 | 86.803 | 66.983 | 159.835 | 1.296 | 13.452 | 0.191 | 1572.038 |
Al | 79.646 | 28.711 | 76.893 | 2.774 | 2.868 | 0.339 | – |
Al Exp. | 79.3 [42] | 29.4 [42] | – | 2.493 [44] | – | 0.323 [44] | ∼960 |
Calculated bulk modulus (GPa), shear modulus (GPa), and elastic anisotropic factors.
Model | BV | BR | GV | GR | AU | A |
---|---|---|---|---|---|---|
CeTi2Al20 | 81.148 | 81.148 | 61.727 | 59.908 | 0.152 | 1.424 |
CeV2Al20 | 76.077 | 76.077 | 63.941 | 63.930 | 0.001 | 1.027 |
CeCr2Al20 | 85.937 | 85.937 | 47.656 | 46.898 | 0.081 | 1.295 |
CeNb2Al20 | 85.393 | 85.393 | 65.474 | 65.388 | 0.007 | 1.077 |
CeTa2Al20 | 86.803 | 86.803 | 67.022 | 66.943 | 0.006 | 1.073 |
Al | 79.646 | 79.646 | 28.972 | 28.449 | 0.092 | 1.318 |

Calculated elastic moduli for different CeT2Al20 intermetallics.
The calculated elastic moduli of CeT2Al20 and pure Al are listed in Tables 3 and 4 together with the experimental values of Al [42], [44]. The larger values of the bulk modulus B and the shear modulus G correspond to better resistance against volume change under hydrostatic pressure and shape deformation by the shearing force. The values of the elastic modulus B follows the order CeTa2Al20 > CeCr2Al20 > CeNb2Al20 > CeTi2Al20 > Al > CeV2Al20 in Table 3, indicating that CeTa2Al20 has the highest resistance against volume deformation. The values of the elastic modulus G follow the order CeTa2Al20 > CeNb2Al20 > CeV2Al20 > CeTi2Al20 > CeCr2Al20 > Al, indicating that CeTa2Al20 has the highest resistance against shape change under external conditions. In comparison to pure Al, the five CeT2Al20 intermetallic compounds have better resistance to shear deformation. The larger moduli G and E correspond to higher hardness of solid materials. The values of G and E of the five intermetallics are obviously larger than those of pure Al, indicating that these ternary intermetallics are harder than pure Al. The hardness was calculated by using the semi-empirical equations (4). The values of hardness are 12.138, 12.827, 7.149, 13.114, 13.452, and 2.868 GPa for CeTi2Al20, CeV2Al20, CeCr2Al20, CeNb2Al20, CeTa2Al20, and Al. Figure 5 displays the elastic moduli E, G, Hv, and C44 for the different CeT2Al20 intermetallics. The elastic moduliE, G, and Hv show the same trend for the different CeT2Al20 intermetallics and pure Al, indicating that the CeT2Al20 intermetallics are obviously harder than Al and that CeTa2Al20 has the highest hardness among these ternary intermetallics.

Calculated B/G ratio and Poisson’s ratio for different CeT2Al20 intermetallics.
The ductile or brittle behaviour influences the high-temperature applications of solid materials. The values of the ratio B/G and Poisson’s ratio v can determine the brittle and ductile behaviour of solid materials. When the values of B/G and v are smaller than 1.75 and 0.26, respectively, solid materials exhibit brittle behaviour; otherwise, they show ductile behaviour. The calculated values of B/G and v are shown in Table 3 and Figure 6. Except for CeCr2Al20, the values of B/G and v for the other CeT2Al20 intermetallics are smaller than the critical values 1.75 and 0.26, indicating that all the CeT2Al20 intermetallics, other than CeCr2Al20, are brittle materials. CeCr2Al20 has larger values of B/G and v, indicating that it is ductile in nature among these intermetallics. The values of the elastic constants (C11–C12) and Young’s modulus are important parameters to estimate the plasticity of solid materials. Generally speaking, if the materials have smaller values of (C11–C12) and E, they have better plasticity [44]. In order to intuitively describe the plastic behaviour, the results for the CeT2Al20 compounds and the experimental values of pure Al [44] are shown in Figure 7. Al has the best plasticity, whereas the five CeT2Al20 compounds have worse plasticity than pure Al. In order to present the brittleness of CeT2Al20 intermetallics, the re-normalised hyperbolic correlations derived by (C12–C44)/E and G/B are shown in Figure 8, including the values of CeT2Al20 and some typical materials, namely Au, Pd, Al, Si, Ge, BN, diamond, and so on [48]. According to the Pugh criterion and the Pettifor criterion, Al and CeCr2Al20 intermetallics are ductile materials, whereas the other four CeT2Al20 intermetallics are all brittle and have high hardness, which is close to that of Si, Ge and BN. Moreover, we also calculated the melting points of CeT2Al20 compounds using the empirical equation [48]

Values of C11–C12 and Young’s modulus for pure Al and CeT2Al20 intermetallics.

Re-normalised hyperbolic correlation derived by (C12–C44)/E and G/B.
The values of the melting point are between 1300 and 1500 K for the five CeT2Al20 intermetallics, as shown in Table 3.
3.3 Anisotropic Properties
The anisotropic properties of solid materials are closely related to the micro-cracks. The elastic anisotropic properties are described by the universal and shear anisotropic indices Au and A. The anisotropic constants are calculated by following equations [48]:
The calculated anisotropic constants are shown in Table 4. When Au and A approach 0 and 1, respectively, crystalline materials show isotropic properties. According to the anisotropic constants, the indices Au and A correspond to the same order: CeTi2Al20 > Al > CeCr2Al20 > CeNb2Al20 > CeTa2Al20 > CeV2Al20, which indicates that CeTi2Al20 has the largest anisotropy and CeV2Al20 the smallest. For CeNb2Al20, CeTa2Al20, and CeV2Al20, the indices Au and A are very close to the characteristic values of 0 and 1, indicating that they are nearly isotropic.
In order to visually describe the anisotropic properties of CeT2Al20 intermetellics, the three-dimensional (3D) surface of the Young’s modulus E was used to further explore their anisotropy. The 3D figures are given by the following equation [49]:

Surface contours of the Young’s modulus for (a) CeTi2Al20, (b) CeV2Al20, (c) CeCr2Al20, (d) CeNb2Al20, (e) CeTa2Al20, and (f) Al.
For crystalline materials, if they possess isotropic properties, their 3D surfaces will be spherical; otherwise, they will be anisotropic. The greater the degree of deviation from a sphere, the stronger the anisotropy. The 3D figures of Young’s modulus are shown in Figure 9. The figures display the degree of deviation, which is small, indicating that these compounds are anisotropic but the anisotropy is small. The 3D figures for CeNb2Al20, CeTa2Al20, and CeV2Al20 are close to a sphere and the degree of deviation from the sphere is relatively small among the six figures, suggesting that they are nearly isotropic. Based on the degree of deviation from the sphere, CeTi2Al20 has the largest anisotropy. In order to further describe the anisotropic properties of the CeT2Al20 intermetellics, the two-dimensional figures of Young’s modulus on different planes are shown in Figure 10. If the contour of the closed curve on the planes is a circle, the crystalline material will have isotropic properties; otherwise, it will have anisotropic properties. In Figure 10, the contours of CeNb2Al20, CeTa2Al20, and CeV2Al20 are close to circles on the different planes, indicating that these intermetallics have elastic properties closer to isotropic.

Projections in the (001) and (110) planes of Al and CeT2Al20 intermetallics.
To further investigate the elastic anisotropy, the acoustic velocities along different directions were calculated by the Brugger equations [50]:
The solid density ρ(g/cm−3) and anisotropic acoustic velocities (km s−1) for CeT2Al20 intermetallics and pure Al are shown in Table 5. According to (8), the elastic constants C11 and C44 determine the longitudinal and transverse acoustic velocities, respectively. The elastic constants C11 are obviously larger than C44 in Table 2, so the longitudinal wave velocities are larger than the transverse ones along the different directions. The calculated results indicate that the acoustic velocities for CeT2Al20 intermetallics are anisotropic. These CeT2Al20 intermetallics have the maximum longitudinal sound velocities along the <111> direction, especially. Furthermore, the difference in the acoustic velocities also implies the elastic anisotropy of CeT2Al20 intermetallics.
Sound velocities along different directions (km s−1) and ρ(g cm−3) for Al and CeT2Al20 intermetallics.
Model | ρ | [100]v1 | [010]vt1 | [001]vt2 | [110]v1 | [111]v1 | |||
---|---|---|---|---|---|---|---|---|---|
CeTi2Al20 | 4.046 | 5.703 | 3.941 | 3.941 | 6.096 | 4.670 | 6.221 | 3.528 | 3.528 |
CeV2Al20 | 4.212 | 6.163 | 3.916 | 3.916 | 6.195 | 5.467 | 6.206 | 3.883 | 3.883 |
CeCr2Al20 | 4.323 | 5.689 | 3.482 | 3.482 | 5.927 | 4.328 | 6.004 | 3.207 | 3.207 |
CeNb2Al20 | 4.481 | 6.138 | 3.878 | 3.878 | 6.225 | 5.284 | 6.254 | 3.785 | 3.785 |
CeTa2Al20 | 5.393 | 5.654 | 3.574 | 3.574 | 5.731 | 4.880 | 5.756 | 3.492 | 3.492 |
Al | 2.723 | 6.416 | 3.432 | 3.432 | 6.636 | 4.227 | 6.275 | 3.144 | 3.144 |
Al [44] | – | 6.325 | 3.220 | 3.220 | 6.469 | 4.128 | 6.157 | 3.023 | 3.023 |

Relationship between the Gibbs free energy and temperature of pure Al and CeT2Al20 intermetallics.

Relationship between the bulk modulus and temperature of CeT2Al20 intermetallics.

Relationship between the thermal expansion and temperature of pure Al and CeT2Al20 intermetallics.

Relationship between the entropy and temperature of pure Al and CeT2Al20.

Relationship between the bulk modulus and temperature of pure Al.

Relationship between the entropy and temperature of pure Al.
3.4 Thermodynamic Properties
For these cubic CeT2Al20 intermetallics, we investigated the thermodynamic properties by calculating their dynamic properties under 11 lattice parameters (from 0.94a to 1.06a) with the Phonopy code. According to the obtained F (V,T) versus V data, we fitted them to the EoS and obtained the thermodynamic quantities under different temperatures. The relationships between the thermodynamic quantities and temperature for CeT2Al20 intermetallics and Al are shown in Figures 11–14. Because the melting point of Al is ∼960 K, the thermodynamic quantities between 0 and 900 K are plotted for pure Al. Figure 11 shows the relationships between the Gibbs free energy and temperature for pure Al and CeT2Al20. In Figure 11, the values of the Gibbs free energy obviously decrease with temperature and become more and more negative, indicating that CeT2Al20 intermetallics satisfy the condition for thermodynamic stability at elevated temperature. In the meantime, the values of the Gibbs free energy of CeT2Al20 become more negative than those of pure Al with the increase in temperature, which indicates that CeT2Al20 intermetallics have better thermal stability than pure Al at high temperatures. Figure 12 shows the relationship between the bulk modulus and temperature of pure Al and CeT2Al20 intermetallics. The values of the bulk modulus become smaller with higher temperature. But the variation in the case of pure Al is larger than that of CeT2Al20 intermetallics, indicating that the five intermetallics have better ability to resist volume change than pure Al. Figure 13 shows the relationship between the thermal expansion coefficients and temperature of pure Al and CeT2Al20. In the low-temperature region, the values of pure Al and CeT2Al20 intermetallics are close to each other. But in the high-temperature region, the values of pure Al are obviously larger than those of other five CeT2Al20 intermetallics at the same temperature, indicating that the pure Al deforms more easily at high temperature. Figure 14 shows the relationship between the entropy and temperature of pure Al and the CeT2Al20 intermetallics. In Figure 14, the values of entropy of pure Al and CeT2Al20 intermetallics increase rapidly and linearly with temperature. In order to prove the correctness of our calculation methods, we compare the theoretical values and experimental ones [51], [52] of the bulk modulus and entropy for pure Al in Figures 15 and 16. The calculated values are very close to the experimental ones, indicating that our calculation methods and results are reliable.
4 Conclusions
In summary, we used the first-principles method to explore the structural stability, elastic properties, brittle or ductile behaviour, anisotropy, dynamical stability, and thermodynamic properties of pure Al and CeT2Al20 intermetallics. The calculated formation enthalpy and phonon frequencies indicate that these intermetallics satisfy the conditions for structural stability. The elastic constants Cij and elastic moduli B, G, E, and Hv indicate that the CeT2Al20 intermetallics possess higher hardness and the better ability to resist deformation than pure Al. The anisotropic constants and acoustic velocities indicate that the CeT2Al20 intermetallics are anisotropic. CeNb2Al20, CeTa2Al20, and CeV2Al20 exhibit show smaller anisotropy of the properties among these intermetallics. The calculated thermal quantities indicate that CeT2Al20 intermetallics have better thermal properties than pure Al at high temperature.
Acknowledgement
This work was supported by the Program for PhD Start-up Fund of the Liaoning province of China (Grant Nos. 201601161 and 20170520055) and the Program for scientific technology plan of the Educational Department of Liaoning province of China (Grant No. LGD2016015).
References
L. Schlapbach and A. Züttel, Nature 414, 353 (2001).10.1038/35104634Suche in Google Scholar
J. Taend, A. Orthacker, and H. Amenitsch, Acta Mater. 117, 43 (2016).10.1016/j.actamat.2016.07.001Suche in Google Scholar
L. B. Yu, J. Wang, F. S. Qu, M. Wang, and W. Wang, J. Alloys Compd. 737, 655 (2018).10.1016/j.jallcom.2017.12.117Suche in Google Scholar
C. Suwanpreecha, P. Pandee, U. Patakham, and C. Limmaneevichitr, Mater. Sci. Eng. A 709, 46 (2018).10.1016/j.msea.2017.10.034Suche in Google Scholar
T. Dorin, M. Ramajayam, J. Lamb, and T. Langanc, Mater. Sci. Eng. A 707, 58 (2018).10.1016/j.msea.2017.09.032Suche in Google Scholar
T. Tian, X. F. Wang, and W. Li, Solid State Commun. 156, 69 (2013).10.1016/j.ssc.2012.10.021Suche in Google Scholar
E. A. Marquis, D. N. Seidman, and D. C. Dunand, Acta Mater. 50, 4021 (2002).10.1016/S1359-6454(02)00201-XSuche in Google Scholar
S. Iwamura and Y. Miura, Acta Mater. 52, 591 (2004).10.1016/j.actamat.2003.09.042Suche in Google Scholar
A. V. Mikhaylovskay, A. G. Mochugovskiy, V. S. Levchenko, N. Yu. Tabachkova, W. Mufalo, et al., Mater. Charact. 139, 30 (2018).10.1016/j.matchar.2018.02.030Suche in Google Scholar
Z. Chen, P. Zhang, D. Chen, Y. Wu, and M. Wang, J. Appl. Phys. 117, 085904 (2015).10.1063/1.4913664Suche in Google Scholar
E. Clouet, A. Barbu, L. La, and G. Martin, Acta Mater. 53, 2313 (2005).10.1016/j.actamat.2005.01.038Suche in Google Scholar
N. Q. Vo, D. C. Dunand, and D. N. Seidman, Acta Mater. 63, 73 (2014).10.1016/j.actamat.2013.10.008Suche in Google Scholar
A. B. Spierings, K. Dawson, T. Heeling, P. J. Uggowitzer, R. Schäublin, et al., Mater. Des. 115, 52 (2017).10.1016/j.matdes.2016.11.040Suche in Google Scholar
J. D. Lin, P. Okle, and D. C. Dunand, Mater. Sci. Eng. A 680, 64 (2017).10.1016/j.msea.2016.10.067Suche in Google Scholar
S. P. Wen, W. Wang, and W. H. Zhao, J. Alloys Compd. 687, 143 (2016).10.1016/j.jallcom.2016.06.045Suche in Google Scholar
I. Halevy, E. Sterer, M. Aizenshtein, G. Kimmel, D. Regev, et al., J. Alloys Compd. 319, 19 (2001).10.1016/S0925-8388(01)00881-7Suche in Google Scholar
A. Yamada, R. Higashinaka, T. D. Matsuda, and Y. Aoki, J. Phys. Soc. Jpn. 87, 033707 (2018).10.7566/JPSJ.87.033707Suche in Google Scholar
A. I. Bram, A. Venkert, and L. Meshi, J. Alloys Comp. 641, 1 (2015).10.1016/j.jallcom.2015.03.247Suche in Google Scholar
K. Wakiya, T. Onimaru, K. T. Matsumoto, and Y. Yamane, J. Phys. Soc. Jpn. 86, 3 (2017).10.7566/JPSJ.86.034707Suche in Google Scholar
M. J. Winiarski, J. C. Griveau, E. Colineau, K. Wochowski, and P. Wisniewski, J. Alloys Compd. 696, 1113 (2017).10.1016/j.jallcom.2016.12.033Suche in Google Scholar
P. C. Liu, Y. J. Xian, X. Wang, Y. T. Zhang, and P. C. Zhang, J. Nucl. Mater. 493, 147 (2017).10.1016/j.jnucmat.2017.05.051Suche in Google Scholar
M. J. Winiarski and T. Klimczuk, J. Solid State Chem. 245, 10 (2017).10.1016/j.jssc.2016.09.029Suche in Google Scholar
C. Moussa, A. Berche, M. Pasturel, J. Barbosa, B. Stepnik, et al., J. Alloys Comp. 691, 893 (2017).10.1016/j.jallcom.2016.08.257Suche in Google Scholar
M. J. Winiarski and T. Klimczuk, Intermetallics 85, 103 (2017).10.1016/j.intermet.2017.02.005Suche in Google Scholar
M. J. Winiarski, B. Wiendlocha, M. Sternik, P. Winiewski, and J. R. O’Brien, Phys. Rev. B. 93, 134507 (2016).10.1103/PhysRevB.93.134507Suche in Google Scholar
A. Sakai and S. Nakatsuji, Phys. Rev. B. 84, 201106 (2011).10.1103/PhysRevB.84.201106Suche in Google Scholar
P. Swatek and D. Kaczorowski, J. Magn. Magn. Mater. 416, 348 (2016).10.1016/j.jmmm.2016.04.086Suche in Google Scholar
M. Tsujimoto, Y. Matsumoto, T. Tomita, A. Sakai, and S. Nakatsuji, Phys. Rev. Lett. 113, 267001 (2014).10.1103/PhysRevLett.113.267001Suche in Google Scholar PubMed
K. R. Kumar, H. S. Nair, R. Christian, A. Thamizhavel, and A. M. Strydom, J. Phys.: Condens. Matter 28, 436002 (2016).10.1088/0953-8984/28/43/436002Suche in Google Scholar PubMed
Y. Verbovytsky, K. Latka, and K. Tomala, J. Alloys Compd. 442, 334 (2007).10.1016/j.jallcom.2006.07.148Suche in Google Scholar
M. J. Kangas, D. C. Schmitt, A. Sakai, S. Nakatsuji, and J. Y. Chan, J. Solid State Chem. 196, 274 (2012).10.1016/j.jssc.2012.06.035Suche in Google Scholar
V. M. T. Thiede, W. Jeitschko, S. Niemann, and T. Ebel, J. Alloys Comp. 267, 23 (1998).10.1016/S0925-8388(97)00532-XSuche in Google Scholar
A. Togo and I. Tanaka, Scr. Mater. 108, 1 (2015).10.1016/j.scriptamat.2015.07.021Suche in Google Scholar
S. Shang, Y. Wang, and Z. K. Liu, Appl. Phys. Lett. 90, 101909 (2007).10.1063/1.2711762Suche in Google Scholar
Y. Pan, Adv. Eng. Mater. 19, 1700099 (2017).10.1002/adem.201700099Suche in Google Scholar
Y. Pan, Mater. Res. Bull. 93, 56 (2017).10.1016/j.materresbull.2017.04.041Suche in Google Scholar
C. Lu, M. S. Miao, and Y. M. Ma, J. Am. Chem. Soc. 135, 14167 (2013).10.1021/ja404854xSuche in Google Scholar PubMed
C. Lu, J. J. Wang, P. Wang, X. X. Xia, Y. Y. Jin, et al., Phys. Chem. Chem. Phys. 19, 1420 (2017).10.1039/C6CP07624FSuche in Google Scholar PubMed
L. P. Ding, P. Shao, F. H. Zhang, C. Lu, L. Ding, et al., Inorg. Chem. 55, 7033 (2016).10.1021/acs.inorgchem.6b00899Suche in Google Scholar PubMed
B. L. Chen, W. G. Sun, X. Y. Kuang, C. Lu, X.X. Xia, et al., Inorg. Chem. 57, 343 (2018).10.1021/acs.inorgchem.7b02585Suche in Google Scholar PubMed
P. H. Dederichs, H. Schober, and D. J. Sellmyer, Electron States and Fermi Surfaces of Alloys, Springer, Berlin, Heidelberg 1981.Suche in Google Scholar
Z. Li and J. S. Tse, Phys. Rev. B 61, 14531 (2000).10.1103/PhysRevB.61.14531Suche in Google Scholar
Z. Mao, W. Chen, D. N. Seidman, and C. Wolverton, Acta Mater. 59, 3012 (2011).10.1016/j.actamat.2011.01.041Suche in Google Scholar
G. Grimvall, Thermo Physical Properties of Materials, North Holland, Amsterdam 1999.Suche in Google Scholar
X. Zhang, Y. Lv, C. Liu, and W. Jiang, Mater. Des. 133, 476 (2017).10.1016/j.matdes.2017.08.014Suche in Google Scholar
S. Chen, Y. Sun, Y. H. Duan, B. Huang, and M. J. Peng, J. Alloys Compd. 630, 202 (2015).10.1016/j.jallcom.2015.01.038Suche in Google Scholar
Y. Lv, X. Zhang, and W. Jiang, Ceram. Int. 44, 128 (2018).10.1016/j.ceramint.2017.09.147Suche in Google Scholar
H. Y. Niu, X. Q. Chen, P. T. Liu, W. W. Xing, X. Y. Cheng, et al., Sci. Rep. 2, 718 (2012).10.1038/srep00718Suche in Google Scholar PubMed PubMed Central
E. Haque and M. A. Hossain, J. Alloys Compd. 748, 117 (2018).10.1016/j.jallcom.2018.03.151Suche in Google Scholar
K. Brugger, J. Appl. Phys. 36, 768 (1965).10.1063/1.1714216Suche in Google Scholar
D. C. Wallace, Ed., Thermodynamics of Crystals, Dover, New York 1972.10.1119/1.1987046Suche in Google Scholar
I. Barin, Thermochemical Date of Pure Substance, Science Press, Beijing 1995.10.1002/9783527619825Suche in Google Scholar
©2018 Walter de Gruyter GmbH, Berlin/Boston
Artikel in diesem Heft
- Frontmatter
- General
- Investigation of Gamma-Ray’s Transmission Geometries for the Measurement of Attenuation Coefficients
- Atomic, Molecular & Chemical Physics
- Theoretical Investigations on the Structural, Electronic and Spectral Properties of VFn (n = 1–7) Clusters
- Dynamical Systems & Nonlinear Phenomena
- Parametric Instability of a Rotating Axially Loaded FG Cylindrical Thin Shell Under Both Axial Disturbances and Thermal Effects
- Generalised Sasa–Satsuma Equation: Densities Approach to New Infinite Hierarchy of Integrable Evolution Equations
- Quantum Theory
- The Schrödinger Equation and Negative Energies
- Gravitational Drift Instability in Quantum Dusty Plasmas
- Hydrodynamics
- Unsteady Peristaltic Transport of a Particle-Fluid Suspension: Application to Oesophageal Swallowing
- Solid State Physics & Materials Science
- First-Principles Investigation of Structural Stability, Mechanical, Anisotropic, and Thermodynamic Properties of CeT2Al20 Intermetallics
Artikel in diesem Heft
- Frontmatter
- General
- Investigation of Gamma-Ray’s Transmission Geometries for the Measurement of Attenuation Coefficients
- Atomic, Molecular & Chemical Physics
- Theoretical Investigations on the Structural, Electronic and Spectral Properties of VFn (n = 1–7) Clusters
- Dynamical Systems & Nonlinear Phenomena
- Parametric Instability of a Rotating Axially Loaded FG Cylindrical Thin Shell Under Both Axial Disturbances and Thermal Effects
- Generalised Sasa–Satsuma Equation: Densities Approach to New Infinite Hierarchy of Integrable Evolution Equations
- Quantum Theory
- The Schrödinger Equation and Negative Energies
- Gravitational Drift Instability in Quantum Dusty Plasmas
- Hydrodynamics
- Unsteady Peristaltic Transport of a Particle-Fluid Suspension: Application to Oesophageal Swallowing
- Solid State Physics & Materials Science
- First-Principles Investigation of Structural Stability, Mechanical, Anisotropic, and Thermodynamic Properties of CeT2Al20 Intermetallics