Next Article in Journal
Mg2FeH6 Synthesis Efficiency Map
Previous Article in Journal
Specific Structural Disorder in an Anion Layer and Its Influence on Conducting Properties of New Crystals of the (BEDT-TTF)4A+[M3+(ox)3]G Family, Where G Is 2-Halopyridine; M Is Cr, Ga; A+ Is [K0.8(H3O)0.2]+
Previous Article in Special Issue
Effect of Sonication Output Power on the Crystal Structure and Magnetism of SrFe12O19 Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Investigations of the Structural, Anisotropic Mechanical, Thermodynamic and Electronic Properties of the AlNi2Ti Compound

State Key Laboratory for Mechanical Behavior of Materials, Xi’an Jiaotong University, Xi’an 710049, China
*
Author to whom correspondence should be addressed.
Crystals 2018, 8(2), 93; https://doi.org/10.3390/cryst8020093
Submission received: 26 December 2017 / Revised: 4 February 2018 / Accepted: 8 February 2018 / Published: 11 February 2018
(This article belongs to the Special Issue Crystal Structure of Magnetic Materials)

Abstract

:
In this paper, the electronic, mechanical and thermodynamic properties of AlNi2Ti are studied by first-principles calculations in order to reveal the influence of AlNi2Ti as an interfacial phase on ZTA (zirconia toughened alumina)/Fe. The results show that AlNi2Ti has relatively high mechanical properties, which will benefit the impact or wear resistance of the ZTA/Fe composite. The values of bulk, shear and Young’s modulus are 164.2, 63.2 and 168.1 GPa respectively, and the hardness of AlNi2Ti (4.4 GPa) is comparable to common ferrous materials. The intrinsic ductile nature and strong metallic bonding character of AlNi2Ti are confirmed by B/G and Poisson’s ratio. AlNi2Ti shows isotropy bulk modulus and anisotropic elasticity in different crystallographic directions. At room temperature, the linear thermal expansion coefficient (LTEC) of AlNi2Ti estimated by quasi-harmonic approximation (QHA) based on Debye model is 10.6 × 10−6 K−1, close to LTECs of zirconia toughened alumina and iron. Therefore, the thermal matching of ZTA/Fe composite with AlNi2Ti interfacial phase can be improved. Other thermodynamic properties including Debye temperature, sound velocity, thermal conductivity and heat capacity, as well as electronic properties, are also calculated.

Graphical Abstract

1. Introduction

AlNi2Ti is one of the Heusler compounds first identified by Heusler in 1903 [1]. The Heusler compounds are a group of ternary intermetallic alloys formed at the stoichiometric composition X2YZ (generally, X and Y are transition metals and Z is a b-subgroup element), with the doubly ordered L21 structure based on the ordered B2 [2]. They have been widely reported to have excellent ferromagnetic, paramagnetic, mechanical and thermoelectric properties [3,4]. AlNi2Ti with L21 structure can provide NiAl or Ni-base and Ti-base alloys with good high-temperature mechanical properties by precipitate strengthening [5,6,7,8,9,10,11,12,13,14,15,16]. Especially the mechanical properties of Fe-Ni-Al-Ti ferritic alloys, such as creep resistance and yield strength, can be significantly enhanced due to the AlNi2Ti precipitates [17,18,19,20,21,22,23,24,25,26]. In addition, AlNi2Ti precipitate phase has also been observed in microstructures of Al-doped Ni-Ti shape memory alloys that have huge potential to be applied as functional material in many areas such as clinical medicine, biotechnology, automation, energy engineering, electronics industry, aeronautics and astronautics, owing to their pseudoelasticity, superplasticity and shape memory properties [27,28,29,30,31,32].
AlNi2Ti is also a common reaction product at the interface between phases. Recently, AlNi2Ti was observed at the interface in zirconia-toughened alumina (ZTA) or Al2O3 reinforced iron matrix composite with an Ni-Ti transition layer that is expected to be applied for wear-resistant material [33,34]. The active elements Ni and Ti can promote the atom diffusion and chemical reaction between ZTA and Fe. Thus, Al2O3 will decompose and diffuse into the Ni-Ti layer, leading to the formation of interfacial phase AlNi2Ti. The mechanical and thermal performance of AlNi2Ti is a key point for improving the interface properties and then wear resistance of composites, especially the thermal expansion coefficient (TEC) that influences the thermal matching of composite interfaces.
By now, a number of research works have been carried out on AlNi2Ti and other Heusler compounds by first-principles calculations. Wen et al. [35] reported the elastic constants and mechanical moduli under different pressures or temperatures of Ni2XAl (X = Sc, Ti, V) compounds. The thermodynamic properties including heat capacity and thermal expansion coefficient as a function of temperature and pressure were demonstrated. They claimed that the influence of pressure on moduli and hardness decreases as the sequence of Ni2ScAl > Ni2TiAl > Ni2VAl, while it has an opposite effect on ductility or anisotropy. Reddy and Kanchana [36] studied the effect of pressure and temperature on the Fermi surface and dynamic properties of a series of compounds Ni2XAl (X = Ti, V, Zr, Nb, Hf and Ta) using density functional theory (DFT). Their calculating results suggested that AlNi2Ti has lower specific heat and entropy but higher internal energy and free energy among the considered compounds [36]. The superconducting property of Ni2NbAl has also been studied and the critical superconducting transition temperature Tc calculated to be 3.1 K [36], which agrees well with the experimental result (Tc around 2.1 K) in previous work [37]. Compared with representative high-pressure superconductors, such as H3S (Tc: 178 K), PH3 (Tc: 81 K) [38] and PtH (Tc: 12.94~20.01 K) [39], Ni2NbAl performs at a relatively low value of Tc. Sahariya and Ahuja [40] presented the energy bands and density of states of AlNi2Ti calculated using the DFT by linear combination of atomic orbitals (LCAO) approach. Zhou et al. [41] investigated the electronic and magnetic properties of AlNi2Ti affected by swap defects and atomic antisite using first-principles calculations. The NiTi(A) antisite was found to be the most probable defect because of the lowest formation energy. Nevertheless, the properties of AlNi2Ti are still short of systematic theoretical study, especially the thermal properties and their anisotropies.
In this paper, a comprehensive and complementary study of the mechanical and thermal properties of AlNi2Ti crystal is conducted by first-principles calculation in order to provide significant theoretical directions for the fabrication and application of composites with interfacial phase AlNi2Ti and to offer a reference for the future research into AlNi2Ti or other Heusler alloys.

2. Calculation Methods

The L21 cubic crystal structure with the F m 3 ¯ m space group of AlNi2Ti is shown in Figure 1. The unit cell of AlNi2Ti has four formulas with a total of 16 atoms.
The first-principles calculations based on DFT were implemented in Cambridge Serial Total Energy Package (CASTEP) code [42,43,44]. The ultra-soft pseudo potentials were employed to represent the interactions between ionic core and valence electrons. The valence electron configurations for Al, Ni and Ti were 3s2 3p1, 3d8 4s2 and 3s2 3p6 3d2 4s2, respectively. A special k-point mesh was used for the numerical integrations in the entire Brillouin zone by setting as 8 × 8 × 8 grid [45]. The Broyden–Fletcher–Goldfarb–Shannon (BFGS) minimization scheme was utilized for the optimization of lattice parameters and atomic positions in order to obtain the ground-state crystal structure. For the purpose of providing more abundant and reliable results, two approaches, i.e., the generalized gradient approximation (GGA) of Perdew Burke Ernzerhof (PBE) and Perdew Wang in 1991 (PW91) approaches, were adopted in exchange–correlation energy calculations [46]. It is capable of comparison with other DFT studies on AlNi2Ti by the same PBE-GGA approach [35,36] and local density approximation (LDA) approach [7,35]. A kinetic energy cut-off value of 500 eV was used for plane wave expansions. The self-consistence convergence of the energy was set to 0.5 × 10−6 eV/atom. The quasi-harmonic approximation (QHA) Debye model was applied to predict the thermal properties of AlNi2Ti compounds.

3. Results and Discussion

3.1. Crystal Parameters and Mechanical Properties

The optimized ground-state structural parameters of AlNi2Ti compounds calculated by the GGA-PBE and GGA-PW91 approaches are shown in Table 1. It can be seen that the lattice parameters at 0 K in this study are close to other theoretical and experimental data [22,35,36,47,48,49,50,51,52,53,54]. The average deviation of results in this work to references for lattice parameters is less than 1%. The chemical stability of the AlNi2Ti compound was evaluated by the cohesive energy Ecoh and formation enthalpy ΔrH, which can be calculated using equations as follows [55].
Ecoh(AlNi2Ti) = Etot(AlNi2Ti) − Eiso(Al) − 2Eiso(Ni) − Eiso(Ti)
ΔrH(AlNi2Ti) = Ecoh(AlNi2Ti) − Ecoh(Al) − 2Ecoh(Ni) − Ecoh(Ti)
where Etot represents the total energy of material and Eiso represents the total energy of a single atom. From the calculated result in Table 1, the negative values of formation enthalpy and cohesive energy indicate the thermodynamic stability of AlNi2Ti.
In order to study the mechanical properties of AlNi2Ti under external stress, the second-order elastic constants cij are calculated by stress versus strain without stress controls. The stress–strain method is conducted using strain modes [ε]T = η[1 0 0 1 0 0], where [ε]T is the transpose of the strain matrix with Voigt notation and η is the magnitude of strain [56]. The maximum strain amplitude is set as 0.003 and the number of steps for each strain is four. Then elastic constants cij can be computed by Hooker’s law, as presented in Table 2. Our results are similar to other calculation data [18,35,36] and the values from the GGA-PW91 method are quite close to those from GGA-PBE. All elastic constants satisfy the Born–Huang stability criterion, so the AlNi2Ti compound has intrinsic stability [57]. The detailed criterion for cubic structure is that an elastically stable crystal should follow restrictions: c44 > 0, c11 > c12, and c11 + 2c12 > 0 [58,59,60]. The value of c11 is much larger than c12 and c44, indicating that it is stiff against uniaxial strain ε1. Then the bulk modulus (B) and shear modulus (G) are computed by Voigt–Reuss–Hill approximation [61]. The Young’s modulus E, Poisson’s ratio v and Vicker’s hardness Hv are estimated by well-known relations E = 9BG/(3B + G), v = (3B − 2G)/(6B + 2G) [62,63,64,65] and Hv = 2(k2G)0.585 − 3, where k = G/B [66]. The results show that the hardness of AlNi2Ti (4.4 GPa) is comparable to some common ferrous materials like pearlite steel (about 2 GPa), martensite steel (about 5.5 GPa) and white cast iron (about 6 GPa). Compared with Ni-Ti binary compounds, the hardness of AlNi2Ti is much higher than NiTi (2.7 or 1.1 GPa with B2 or B19’ structure respectively) and NiTi2 (1.6 GPa), and the mechanical moduli of AlNi2Ti are also higher than NiTi and NiTi2, but lower than Ni3Ti [55]. The Poisson’s ratio of AlNi2Ti is determined as 0.32. Since the Poisson’s ratio is normally 0.25 for covalent materials and 0.3 for metallic materials, our result indicates that AlNi2Ti has a strong metallic bonding character. Besides, the value of B/G is a criterion for justifying the ductility of the material; for tough materials, B/G > 1.75, and for brittle materials, B/G < 1.75. The B/G value (2.6) indicates that AlNi2Ti has relatively higher toughness and ductility than common ceramics, such as WC (B/G = 1.37 [67]), SiO2 (1.39 [68]), TiC (1.41 [69,70]), Al2O3 (1.52 [71]) and ZrO2 (2.3 [71]). Therefore, the existence of AlNi2Ti at the ZTA/Fe interface will benefit the mechanical properties of the composite with considerable certainty. Our results are close to other calculated data, except for hardness, which was estimated by another empirical equation in reference [35].
The Helmholtz free energy is calculated by the following equation to investigate the stability of AlNi2Ti under elevated temperature [45]:
E(V, T) = Egs(V) + Evib(V, T) + Eele(V, T),
where Egs refers to the ground state total energy; Evib refers to vibrational free energy which can be given by QHA approximation based on the Debye model; Eele refers to the electron thermal excitations at finite temperature. Then the equilibrium volume and bulk modulus at the certain temperature can be obtained from isothermal curves (E(V, T)-V) by fitting the Birch–Murnaghan equation of state as follows [72,73,74]:
E ( V ) = B 0 V 0 B 0 [ 1 B 0 1 ( V V 0 ) ( 1 B 0 ) + V V 0 + B 0 1 B 0 ] + E 0 ,
Here E(V) represents the Helmholtz free energy of crystal with a volume of V; E0, B0 and V0 are the energy, bulk modulus and cell volume, respectively, under 0 K; B0 refers to the first-order derivative of bulk modulus versus pressure under 0 K. Then the computed lattice parameters and bulk moduli as the function of temperature are shown in Figure 2. It can be seen that with the rise of temperature from 0 to 1500 K, the lattice parameter increases from 5.91 to 6.01 Å while the bulk modulus declines from 162.95 to 138.05 GPa. The result indicates that AlNi2Ti is not particularly stable under elevated temperatures and there might be a certain amount of volume expansion when the temperature rises.
In addition, the compressibility calculated from the ratio of volume (V) under a certain pressure to original volume (V0) without pressing is adopted to characterize the stability under high pressure and plasticity. Figure 3 shows the computed results of AlNi2Ti as well as some ceramic and metal materials for comparison. From Figure 3, under the same pressure, the volume change of AlNi2Ti is larger than TiO, TiO2, Fe, α-Al2O3 and diamond, so the intermetallic compound AlNi2Ti performs the best plasticity and deformation ability.

3.2. Anisotropy of Mechanical Moduli

The elastic anisotropy of AlNi2Ti with cubic structure is investigated by computing the values of bulk and Young’s moduli under different directions, as in the following equations [75]:
B = 1 / ( s 11 + 2 s 12 ) ,
E = 1 / [ s 11 2 ( s 11 s 12 1 2 s 44 ) ( l 1 2 l 2 2 + l 2 2 l 3 2 + l 3 2 l 1 2 ) ] ,
Here sij represents the elastic compliance matrix, which is the inverse matrix of cij. l1, l2 and l3 are the directional cosines. In spherical coordinates, l1, l2 and l3 are computed by l1 = sinθcosφ, l2 = sinθsinφ, and l3 = cosθ respectively. Then the obtained three-dimensional contour plots of elastic moduli are shown in Figure 4, revealing that the bulk modulus of AlNi2Ti shows spherical morphology while the Young’s modulus shows relatively stronger anisotropy. From the planar projections of bulk and Young’s moduli on [001] and [110] crystallographic planes in Figure 4, the anisotropy of the Young’s modulus on both [001] and [110] planes is apparent. The bulk and Young’s moduli on principal crystallographic directions can be calculated by
B[100] = 1/(s11 + s12 + s13), B[010] = 1/(s12 + s22 + s23), B[001] = 1/(s13 + s23 + s33),
E[100] = 1/s11, E[010] = 1/s22, E[001] = 1/s33,
The results indicate that the bulk modulus of AlNi2Ti is B[100] = B[010] = 492.4 GPa, which shows obvious isotropy. As for the Young’s modulus, the maximum and minimum values on the [001] plane are E[110] = 173.4 GPa and E[100] = 92.7 GPa; the maximum and minimum values on the [110] plane are E[111] = 244.4 GPa and E[001] = 92.7 GPa, respectively.
Three shear anisotropic factors (A1, A2 and A3) for cubic crystal are calculated to characterize the elastic anisotropy, which were defined by Ravindran et al. as A1 = A2 = A3 = 2c44/(c11c12) [76]. As in Table 2, the computed values of A1, A2 and A3 were 2.96 by the GGA-PBE method and 2.88 by the GGA-PW91 method, demonstrating that the shear modulus of AlNi2Ti has a strong directional dependence, because the values of A1, A2 and A3 should be one for an isotropic crystal [77]. Besides, Ranganathan and Ostoja-Starzewski proposed another universal anisotropy index AU determined by [78]:
AU = 5(GV/GR) + (BV/BR) − 6,
The elastic anisotropy can also be evaluated by the percent anisotropies AB and AG in compression and shear modes as follows [79]:
AB = (BVBR)/(BV + BR),
AG = (GVGR)/(GV + GR),
The results in Table 2 show that our results are close to the data calculated by others. The value of AU over zero implies that AlNi2Ti shows the universal anisotropy of mechanical moduli; AB has zero value while the value of AG is slightly larger than zero, so AlNi2Ti has isotropic bulk modulus and anisotropic shear modulus [77], which agrees well with the results from the Voigt and Reuss approximations.

3.3. Sound Velocities and Thermal Conductivity

The sound velocity in a crystal is an important parameter for thermodynamic properties. For cubic structure, the sound velocities of both longitudinal and transverse waves on principal crystallographic directions should be evaluated by the procedure of Braggar using elastic constants, as in the relations below [80,81,82,83]:
for   [ 100 ] v l = c 11 / ρ ,   [ 010 ] v t 1 = [ 001 ] v t 2 = c 44 / ρ ,
for   [ 110 ] v l = ( c 11 + c 12 + 2 c 44 ) / 2 ρ ,   [ 1 1 ¯ 0 ] v t 1 = ( c 11 c 12 ) / ρ ,   [ 001 ] v t 2 = c 44 / ρ ,
for   [ 111 ] v l = ( c 11 + 2 c 12 + 4 c 44 ) / 3 ρ ,   [ 11 2 ¯ ] v t 1 = v t 2 = ( c 11 c 12 + c 44 ) / 3 ρ ,
Here ρ is the theoretical density of the compound; vl is the longitudinal sound velocity; vt1 and vt2 are the first and the second transverse mode, respectively. Obviously, c11 determines the longitudinal sound velocity along [100] direction, while c44 corresponds to the transverse modes. The sound velocities in the [110] and [111] directions are related to all values of c11, c12 and c44. The anisotropy of sound velocities implies the elastic anisotropy of crystal. The calculated results are tabulated in Table 3. For comparison, the sound velocities of the Ni3Ti intermetallic compound with a hexagonal structure are also calculated. It is indicated that the longitudinal sound velocities along all directions of Ni3Ti are higher than AlNi2Ti.
Besides, the total transverse, longitudinal and mean sound velocities (vt, vl and vm) can be calculated from density ρ, bulk and shear moduli (B and G), as following equations [84,85].
v t = G ρ ,
v l = ( B + 4 3 G ) ρ ,
v m = [ 1 3 ( 2 v t 3 + 1 v l 3 ) ] 1 / 3 ,
In Table 3, the velocities of AlNi2Ti are slightly lower than Ni3Ti. Once the mean sound velocities are obtained, the Debye temperature (ΘD) can be estimated by [86]:
Θ D = h k B [ 3 n 4 π ( N A ρ M ) ] 1 / 3 v m ,
where h is the Planck’s constant; kB is the Boltzmann constant; NA is the Avagadro’s constant; n is the number of atoms per formula; and M is the molecular weight. The calculated Debye temperature values of AlNi2Ti and Ni3Ti are shown in Table 4. Our results are in good agreement with both the theoretical values and experimental data [35,87]. The Debye temperature of Ni3Ti is higher than that of AlNi2Ti, so the inter-atomic force in Ni3Ti is correspondingly stronger than AlNi2Ti. This may be the reason why the moduli of Ni3Ti are relatively larger than AlNi2Ti.
Thermal conductivity characterizes the thermal transportation behavior of material. In this study, the minimum thermal conductivity (kmin) of AlNi2Ti as well as Ni3Ti for comparison is estimated by Clarke’s model and Cahill’s model as follows [88,89]:
in Clark’s model,
k min = 0.87 k B M ¯ 2 / 3 E 1 / 2 ρ 1 / 6 ,
in Cahill’s model,
k min = k 2.48 P 2 / 3 ( v l + 2 v t ) ,
where kB is the Boltzmann constant; M ¯ = M / ( m · N A ) represents the average mass per atom; E is the Young’s modulus; ρ is the density; P is the density of number of atoms per volume; and vl and vt refer to the longitudinal and transverse sound velocities, respectively. The results in Table 4 show that the minimum thermal conductivities of AlNi2Ti predicted by Clark’s model (1.14 W·m−1 K−1) are slightly smaller than Cahill’s model (1.29 W·m−1 K−1). The minimum thermal conductivities of AlNi2Ti by both models are close to that of Ni3Ti, while the values of AlNi2Ti are relatively smaller, probably due to the higher modulus of Ni3Ti. The calculated thermal conductivities are smaller than the experimental value [2], because the experimental measurement is conducted using AlNi2Ti alloy rather than single phase. Besides, the anisotropy of minimum thermal conductivity estimated by Clark’s model is consistent with that by Cahill’s model to some degree. The kmin has the same value in the [100], [010] and [001] directions by both models. However, the largest kmin estimated within Clark’s model is 1.37 W·m−1 K−1 in [111] direction; as for Cahill’s model, the largest kmin (1.40 W·m−1 K−1) is in [110] direction.

3.4. Thermal Expansion Coefficient and Heat Capacity

The Helmholtz free energy E(V, T)-V curves under different temperatures are obtained and presented in Figure 5. Then the equilibrium volumes at each temperature can be computed by fitting the Birch–Murnaghan equation of state. The volume expansions (V/V0) under various temperatures are shown in Figure 6. It can be seen that the volume expansion of AlNi2Ti increases with temperature and is greater than that of Ni-Ti binary compounds, which were investigated in reference [55].
Then the volumetric TEC λ can be calculated by λ = (∂V/∂T)/V. Moreover, the linear TEC α of cubic AlNi2Ti determined as α = λ/3 is near the LTEC of Ni3Ti, and higher than other Ni-Ti binary compounds, as shown in Figure 7. With the temperature rising, the LTEC of AlNi2Ti increases rapidly first and then shows nearly linear growth with relatively slow increments above 200 K. At room temperature (300 K approximately), the LTECs of AlNi2Ti and Ni-Ti binary compounds are AlNi2Ti: 10.6 × 10−6 K−1, Ni3Ti: 8.92 × 10−6 K−1, B2_NiTi: 7.50 × 10−6 K−1, NiTi2: 6.31 × 10−6 K−1, respectively, so AlNi2Ti has the largest LTEC. However, the LTEC of Ni3Ti increases faster and exceeds the LTEC of AlNi2Ti under elevated temperatures above 1200 K. Since AlNi2Ti is considered one of the interfacial phases in ZTA-reinforced iron matrix composite, the thermal matching property of the ZTA/Fe interface should be improved as much as possible. In this study, the LTEC of AlNi2Ti is quite near to ZTA (7.0–9.5 × 10−6 K−1 [90]) and Fe (9.2–16.9 × 10−6 K−1 [91]), so the ZTA/Fe composite with AlNi2Ti at the interface will have excellent thermal shock resistance.
Figure 7 also shows the LTEC of another Heusler compound AlFe2Ti (14.5 × 10−6 K−1 [52]), which is higher than our result. In addition, the specific heats at constant pressure (Cp) and constant volume (Cv) are studied and their values follow the equation
CpCv = λ2V(T)TB,
where λ is the volumetric thermal expansion coefficient; V(T) and B are the cell volume and crystal bulk modulus at temperature T. The calculated results for specific heats are shown in Figure 8, indicating that our results agree well with the measured Cp of AlNi2Ti in experiment from 102 to 606.1 K [92]. At low temperature, Cp and Cv both increase rapidly with the temperature rising and the increment speed decreases substantially at higher temperatures. The high temperature limit of Cv tends to approaching the classic Dulong–Petit limit, and Cp keeps growing due to the work done by the volumetric expansion.
In order to explore the low-temperature heat capacity of AlNi2Ti with metallic features, Cp can be estimated from the calculations of the electronic structures and elastic constants, which is given by [67].
C p ( T ) = γ T + β T 3 ,  
γ = 1 3 π 2 k B 2 D f ,
β = 12 π 4 R n 5 Θ D 3 ,
where γ and β are the coefficients of electronic and lattice heat, respectively; Df is the density of state (DOS) value at the Fermi level; R is the molar gas constant and n is the total number of atoms per formula unit. It is worth mentioning that ΘD can only describe the temperature-dependence of Cp when the temperature is lower than ΘD/10 [93]. Therefore, the values of Df, γ and β are shown in Table 5 and the Cp-T curve of AlNi2Ti and Ni-Ti binary compounds for comparison from 0 to 45 K is obtained in Figure 9.
From the illustration at the top left corner of Figure 9, it can be seen that at temperatures up to around 5.25 K, the values of Cp have the following order: NiTi2 > B2_NiTi > AlNi2Ti ≈ B19’_NiTi > Ni3Ti, which is the same as the sequence of γ, indicating that the main contribution to Cp is the excitation of electrons at first. The smallest value of γ is from Ni3Ti (γ = 0.267 mJ/(K2·mol)), because it has the smallest DOS value at the Fermi level and thus the weakest metallic nature among these compounds. At the temperature from 5.25 to 45 K, and contributions from phonon excitations (β) have to be taken into account, so the various growth rates result in a different order of Cp: NiTi2 > AlNi2Ti ≈ B19’_NiTi > B2_NiTi ≈ Ni3Ti, the same sequence as β. The calculated results of Cp in Table 5 give the examples below and above 5.25 K, which can demonstrate the distinction clearly. Therefore, it is concluded that the Cp curve is dominated by the heat capacity from electrons at first and when the temperature further rises the heat capacity is mainly determined by phonon excitation.

3.5. Formatting of Mathematical Components

The electronic properties of AlNi2Ti are investigated by calculating the total density of states (TDOS) and partial density of states (PDOS). As shown in Figure 10, there are two sharp peaks which arise from the Ti-3s and 3p states respectively in the lower energy part of the DOS curve. They are localized and have little effect on chemical bonding. Around the Fermi level, the DOS is mainly composed of three wide peaks. The resonance peak I appearing in the DOS is caused by the hybridization of Al-p and Ni-d orbitals. There exists a hollow between peaks II and III, which reflects the splitting of bonding and anti-bonding states. The bonding states (peak II) mainly consists of Al-p and Ni-d states, while the anti-bonding states above the Fermi level are a conduction band composed of 3d bands of Ti, which form peak III. Obviously, the metallic nature of AlNi2Ti can be determined by a nonzero DOS value at the Fermi level (3.81 electrons/eV·f.u.), which provides this compound with an excellent ductile character.

4. Conclusions

The mechanical and thermodynamic properties of AlNi2Ti were investigated by first-principles calculations based on DFT. The mechanical moduli of AlNi2Ti were computed from the independent elastic constants. The computed values of bulk, shear and Young’s modulus were 164.2, 63.2 and 168.1 GPa, respectively, indicating that it has relatively high mechanical properties. The hardness of AlNi2Ti (4.4 GPa) is comparable to common ferrous materials, like martensite steel and white cast iron. Compared with Ni-Ti binary compounds, the hardness and moduli of AlNi2Ti are much higher than NiTi and NiTi2, but lower than Ni3Ti. The intrinsic ductile nature and strong metallic bonding character of AlNi2Ti are confirmed by B/G and Poisson’s ratio. Besides, AlNi2Ti shows the isotropy bulk modulus and anisotropic elasticity in different crystallographic directions. Therefore, the existence of AlNi2Ti at the ZTA/Fe interface will probably benefit the mechanical properties of the composite. At room temperature, the linear thermal expansion coefficient (LTEC) of AlNi2Ti, estimated by quasi-harmonic approximation based on the Debye model, was 10.6 × 10−6 K−1, quite close to LTECs of ZTA and iron, so the thermal shock resistance of the ZTA/Fe composite with an AlNi2Ti interfacial phase can be improved. Other thermodynamic properties including Debye temperature, sound velocity, thermal conductivity and heat capacity were also calculated and compared with Ni-Ti binary compounds. The electronic properties of AlNi2Ti imply its excellent ductile character.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (51501139); the Science and Technology Project of Guangdong Province in China (2015B010122003, 2015B090926009); and the Science and Technology Project of Guangzhou City in China (201604046009).

Author Contributions

Yefei Li conceived the total investigation and performed the calculations; Shuli Tang analyzed the data and wrote the paper; Yimin Gao supervised the whole work; Qiaoling Zheng contributed the device; Zhiwei Liu and Xiangyi Ren revised the paper.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Heusler, F. Ueber magnetische manganlegierungen (on magnetic manganese alloys). Verhandlungen der Deutschen Physikalischen Gesellschaft 1903, 5, 219. (In German) [Google Scholar]
  2. Nakata, J.; Terada, Y.; Takizawa, S.; Ohkubo, K.; Mohri, T.; Suzuki, T. Thermal conductivity in X2YZ heusler type intermetallic compounds. Mater. Trans. JIM 1996, 37, 442–447. [Google Scholar] [CrossRef]
  3. Oxley, D.P.; Tebble, R.S.; Williams, K.C. Heusler alloys. J. Appl. Phys. 1963, 34, 1362–1363. [Google Scholar] [CrossRef]
  4. Bradley, A.J.; Rogers, J.W. The crystal structure of the Heusler alloys. Proc. R. Soc. 1934, A144, 340–359. [Google Scholar] [CrossRef]
  5. Wilson, A.W.; Howe, J.M. Effect of alloying additions on beta’ precipitation in NiAl-Ti base alloys. Acta Mater. 2001, 49, 2653–2660. [Google Scholar] [CrossRef]
  6. Koizumi, Y.; Ro, Y.; Nakazawa, S.; Harada, H. NiTi-base intermetallic alloys strengthened by Al substitution. Mater. Sci. Eng. A 1997, 223, 36–41. [Google Scholar] [CrossRef]
  7. Lin, W.; Freeman, A.J. Cohesive properties and electronic-structure of Heusler L21-phase compounds Ni2XAl (X = Ti, V, Zr, Nb, Hf and Ta). Phys. Rev. B 1992, 45, 61–68. [Google Scholar] [CrossRef]
  8. Ramos, A.S.; Vieira, M.T.; Simoes, S.; Viana, F.; Vieira, M.F. Reaction-assisted diffusion bonding of advanced materials. Defect. Diffus. Forum 2010, 297–301, 972–977. [Google Scholar] [CrossRef]
  9. Bozzolo, G.; Khalil, J.; Bartow, M.R.; Noebe, D. Atomistic modeling of ternary and quaternary ordered intermetallic alloys. In Materials Research Society Symposium Proceedings Vol. 646, Proceedings of the High-Temperature Ordered Intermetallic Alloys IX. Symposium, Boston, MA, USA, 27–29 November 2000; Schneibel, J.H., Hemker, K.J., Noebe, R.D., Hanada, S., Sauthoff, G., Eds.; Materials Research Society: Warrendale, PA, USA, 2001; p. N6.2.1-8. [Google Scholar]
  10. Bozzolo, G.; Noebe, R.B.; Ferrante, J. BFS simulation and experimental analysis of the effect of Ti additions on the structure of NiAl. J. Comput. Aided Mater. Des. 1999, 6, 33–68. [Google Scholar] [CrossRef]
  11. Bozzolo, G.; Noebe, R.B.; Ferrante, J.; Garg, A. Atomistic simulations of alloying additions to NiAl. Mater. Sci. Eng. A 1997, 239–240, 769–776. [Google Scholar] [CrossRef]
  12. Gale, W.F.; Abdo, Z.A.M. Microstructural development in cast alloys based on the β-NiAl–β′-Ni2AlTi–γ′-Ni3Al–α-Cr system. J. Mater. Sci. 1998, 33, 2299–2304. [Google Scholar] [CrossRef]
  13. Field, R.D.; Darolia, R.; Lahrman, D.F. Precipitation in NiAl/Ni2AlTi alloys. Scr. Metall. 1989, 23, 1469–1474. [Google Scholar] [CrossRef]
  14. Cahn, R.W. Nickel Alumina-Titanium Alloy Contg. Epitaxially Related Phases-Having Combined Strength, High Temp. Creep Resistance and Room Temp. Ductility or Plastic Deformability, for Gas Turbine Components. Patent WO9209712-A1, 11 June 1992. [Google Scholar]
  15. Matano, T.; Kimura, Y.; Miura, S.; Mishima, Y. Microstructure and mechanical properties of the L12/L21 two-phase alloys in the quaternary Co-Al-Ni-Ti system. In High-Temperature Ordered Intermetallic Alloys VI. Symposium, Vol. 2, Proceedings of the High-Temperature Ordered Intermetallic Alloys VI. Symposium, Boston, MA, USA, 28 November–1 December 1994; Horton, J., Baker, I., Hanada, S., Noebe, R.D., Schwartz, D.S., Eds.; Materials Research Society: Pittsburgh, PA, USA, 1995; pp. 1377–1382. [Google Scholar]
  16. Mishima, Y.; Lee, E.H.; Liu, C.T. Microstructure, phase constitution and tensile properties of Co-Ni-Ti-Al base multiphase intermetallic alloys. Mater. Trans. 1995, 36, 1031–1040. [Google Scholar] [CrossRef]
  17. Song, G.; Sun, Z.Q.; Poplawsky, J.D.; Xu, X.D.; Chen, M.W.; Liaw, P.K. Primary and secondary precipitates in a hierarchical-precipitate-strengthened ferritic alloy. J. Alloys Compd. 2017, 706, 584–588. [Google Scholar] [CrossRef]
  18. Song, G.; Sun, Z.Q.; Li, L.; Clausen, B.; Zhang, S.Y.; Gao, Y.F.; Liaw, P.K. High temperature deformation mechanism in hierarchical and single precipitate strengthened ferritic alloys by in situ neutron diffraction studies. Sci. Rep. 2017, 7. [Google Scholar] [CrossRef] [PubMed]
  19. Song, G.; Sun, Z.Q.; Clausen, B.; Liaw, P.K. Microstructural characteristics of a Ni2TiAl-precipitate-strengthened ferritic alloy. J. Alloys Compd. 2017, 693, 921–928. [Google Scholar] [CrossRef]
  20. Song, G.L.; Sun, Z.Q.; Poplawsky, J.D.; Gao, Y.F.; Liaw, P.K. Microstructural evolution of single Ni2TiAl or hierarchical NiAl/Ni2TiAl precipitates in Fe-Ni-Al-Cr-Ti ferritic alloys during thermal treatment for elevated-temperature applications. Acta Mater. 2017, 127, 1–16. [Google Scholar] [CrossRef]
  21. Song, G.; Sun, Z.Q.; Li, L.; Xu, X.D.; Rawlings, M.; Liebscher, C.H.; Clausen, B.; Poplawsky, J.; Leonard, D.N.; Huang, S.Y. Ferritic alloys with extreme creep resistance via coherent hierarchical precipitates. Sci. Rep. 2015, 5. [Google Scholar] [CrossRef] [PubMed]
  22. Qian, F.; Sharp, J.; Rainforth, W.M. Microstructural evolution of Mn-based maraging steels and their influences on mechanical properties. Mater. Sci. Eng. A 2016, 674, 286–298. [Google Scholar] [CrossRef]
  23. Qian, F.; Sharp, J.; Rainforth, W.M. Characterisation of L21-ordered Ni2TiAl precipitates in Fe-Mn maraging steel. Mater. Charact. 2016, 118, 199–205. [Google Scholar] [CrossRef]
  24. Liebscher, C.H.; Radmilovic, V.R.; Dahmen, U.; Vo, N.Q.; Dunand, D.C.; Asta, M.; Ghosh, G. A hierarchical microstructure due to chemical ordering in the bcc lattice: Early stages of formation in a ferritic Fe-Al-Cr-Ni-Ti alloy. Acta Mater. 2015, 92, 220–232. [Google Scholar] [CrossRef]
  25. Sun, Z.; Liebscher, C.H.; Huang, S.; Teng, Z.; Song, G.; Wang, G.; Asta, M.; Rawlings, M.; Fine, M.E.; Liaw, P.K. New design aspects of creep-resistant NiAl-strengthened ferritic alloys. Scr. Mater. 2013, 68, 384–388. [Google Scholar] [CrossRef]
  26. Raqlings, M.J.S.; Liebscher, C.H.; Asta, M.; Dunand, D.C. Effect of titanium additions upon microstructure and properties of precipitation-strengthened Fe-Ni-Al-Cr ferritic alloys. Acta Mater. 2017, 128, 103–112. [Google Scholar] [CrossRef]
  27. Jung, J.; Ghosh, G.; Isheim, D.G.; Olson, B. Design of nanodispersion strengthened TiNi-base shape memory alloys. In Proceedings of the International Conference on Shape Memory and Superelastic Technologies, Pacific Grove, CA, USA, 5–8 May 2003; Pelton, A.R., Duerig, T., Eds.; SMST Society Inc.: Menlo Park, CA, USA, 2004; pp. 23–32. [Google Scholar]
  28. Hsu, D.H.D.; Hornbuckle, B.C.; Valderrama, B.; Barrie, F.; Henderson, H.B.; Thompson, G.B.; Manuel, M.V. The effect of aluminum additions on the thermal, microstructural, and mechanical behavior of NiTiHf shape memory alloys. J. Alloys Compd. 2015, 638, 67–76. [Google Scholar] [CrossRef]
  29. Shi, J.; Gao, Z.H.; Hu, K.; Pan, G.J.; Wei, M.Z.; Xu, L.J.; Meng, X.K. High pseudoelasticity of nanoscale L21 phase-Ni43Ti38Al19 thin films. Mater. Lett. 2014, 115, 79–81. [Google Scholar] [CrossRef]
  30. Peters, M.A.; Botton, G.A.; Humphreys, C.J. The precipitation of β’ Ni2TiAl from Al-doped β Ni-Ti alloys. In Institute of Physics Conference Series Vol. 147: Electron Microscopy and Analysis 1995, Proceedings of the Institute of Physics Electron Microscopy and Analysis Group Conference, Birmingham, England, 12–15 September 1995; Cherns, D., Ed.; IOP Publishing Ltd.: Bristol, UK, 1995; pp. 451–454. [Google Scholar]
  31. Jung, J.; Ghosh, G.; Isheim, D.; Olson, G.B. Precipitation of heusler phase (Ni2TiAl) from B2-TiNi in Ni-Ti-Al and Ni-Ti-Al-X (X = Hf, Zr) alloys. Metall. Mater. Trans. A 2003, 34, 1221–1235. [Google Scholar] [CrossRef]
  32. Jung, J.; Ghosh, G.; Olson, G.B. A comparative study of precipitation behavior of Heusler phase (Ni2TiAl) from B2-TiNi in Ni-Ti-Al and Ni-Ti-Al-X (X = Hf, Pd, Pt, Zr) alloys. Acta Mater. 2003, 51, 6341–6357. [Google Scholar] [CrossRef]
  33. Tang, S.L.; Gao, Y.M.; Li, Y.F.; Zheng, Q.L. Preparation and interface investigation of iron matrix composite reinforced by alumina particles sintered with Ti and Ni. Adv. Eng. Mater. 2016, 18, 1913–1920. [Google Scholar] [CrossRef]
  34. Tang, S.L.; Gao, Y.M.; Li, Y.F. Investigation on wear behavior of coated ZTA particles reinforced iron matrix composite. In Proceedings of the 3rd China International Congress on Composite Materials, Hangzhou, China, 21–23 October 2017. (In Chinese). [Google Scholar]
  35. Wen, Z.Q.; Zhao, Y.H.; Hou, H.; Wang, B.; Han, P.D. The mechanical and thermodynamic properties of Heusler compunds Ni2XAl (X = Sc, Ti, V) under pressure and temperature: A first-principles study. Mater. Des. 2017, 114, 398–403. [Google Scholar] [CrossRef]
  36. Reddy, P.V.S.; Kanchana, V. Ab initio study of Fermi surface and dynamical properties of Ni2XAl (X = Ti, V, Zr, Nb, Hf and Ta). J. Alloys Compd. 2014, 616, 527–534. [Google Scholar] [CrossRef]
  37. Waki, S.; Yamaguchi, Y.; Mitsugi, K. Superconductivity of Ni2NbX (X = Al, Ga and Sn). J. Phys. Soc. Jpn. 1985, 54, 1673–1676. [Google Scholar] [CrossRef]
  38. Durajski, A.P. Quantitative analysis of nonadiabatic effects in dense H3S and PH3 superconductors. Sci. Rep. 2016, 6. [Google Scholar] [CrossRef] [PubMed]
  39. Szczȩśniak, D.; Zemła, T.P. On the high-pressure superconducting phase in platinum hydride. Supercond. Sci. Technol. 2015, 28, 085018. [Google Scholar] [CrossRef]
  40. Sahariya, J.; Ahuja, B.L. Electronic structure of Ni2TiAl: Theoretical aspects and Compton scattering measurement. Physica B 2012, 407, 4182–4185. [Google Scholar] [CrossRef]
  41. Zhou, Y.; Chen, Y.; Feng, Y.; Yuan, H.K.; Chen, H. First-principles study on the effect of defects on the electronic and magnetic properties of the Ti2NiAl inverse Heusler alloy. Eur. Phys. J. B 2014, 87, 290. [Google Scholar] [CrossRef]
  42. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulations: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  43. Mattsson, A.E.; Schultz, P.A.; Desjarlais, M.P.; Mattsson, T.R.; Leung, K. Designing meaningful density functional theory calculations in materials science—A primer. Modell. Simul. Mater. Sci. Eng. 2005, 13, R1–R31. [Google Scholar] [CrossRef]
  44. Cao, L.Z.; Shen, J.; Chen, N.X. Theoretical study of the phase stability and site preference for R3(Fe, T)29 (R = Nd, Sm; T = V, Ti, Cr, Cu, Nb, Mo, Ag). J. Alloys Compd. 2002, 336, 18–28. [Google Scholar] [CrossRef]
  45. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  46. Perdew, J.P.; Burke, K.; Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B 1996, 54, 16533–16539. [Google Scholar] [CrossRef]
  47. Debski, A.; Gasior, W.; Sypien, A.; Goral, A. Enthalpy of formation of intermetallic phases from Al-Ni and Al-Ni-Ti systems. Intermetallics 2013, 42, 92–98. [Google Scholar] [CrossRef]
  48. Huneau, B.; Rogl, P.; Zeng, K.; Schmid-Fetzer, R.; Bohn, M.; Bauer, J. The ternary system Al-Ni-Ti Part I: Isothermal section at 900 °C; Experimental investigation and thermodynamic calculation. Intermetallics 1999, 7, 1337–1345. [Google Scholar] [CrossRef]
  49. Yin, M.; Nash, P.; Chen, W.; Chen, S. Standard enthalpies of formation of selected Ni2YZ Heusler compounds. J. Alloys Compd. 2016, 660, 258–265. [Google Scholar] [CrossRef]
  50. Pearson, W.B. Handbook of Lattice Spacings and Structures of Metals; Pergamon: New York, NY, USA, 1967; Volume 2. [Google Scholar]
  51. Jiang, D.T.; Guo, J.T.; Shi, C.X.; Lin, D.L. Microstructure and compressive properties of in situ composite (Ni-40Al-10Ti)-(0, 20%)TiC. J. Mater. Sci. Lett. 2000, 19, 115–117. [Google Scholar] [CrossRef]
  52. Yan, X.L.; Grytsiv, A.; Rogl, P.; Pomjakushin, V.; Palm, M. The Heusler phase Ti25(Fe50-xNix)Al25 (0 ≤ x ≤ 50); structure and constitution. J. Phase Equilib. Diffus. 2008, 29, 500–508. [Google Scholar] [CrossRef]
  53. Sridharan, S.; Nowotny, H.; Wayne, S.F. Investigations within the quaternary system titanium—nickel—alumina—carbon. Monatsh. Chem. 1983, 114, 127–135. [Google Scholar] [CrossRef]
  54. Hu, R.X.; Nash, P.; Chen, Q. Enthalpy of formation in the Al-Ni-Ti system. J. Phase Equilib. Diffus. 2009, 30, 559–563. [Google Scholar] [CrossRef]
  55. Li, Y.F.; Tang, S.L.; Gao, Y.M.; Ma, S.Q.; Zheng, Q.L.; Cheng, Y.H. Mechanical and thermodynamic properties of intermetallic compounds in the Ni-Ti system. Int. J. Mod. Phys. B 2017, 31. [Google Scholar] [CrossRef]
  56. Zhao, J.J.; Winey, J.M.; Gupta, Y.M. First-principles calculations of second- and third-order elastic constants for single crystals of arbitrary symmetry. Phys. Rev. B 2007, 75. [Google Scholar] [CrossRef]
  57. Patil, S.K.R.; Khare, S.V.; Tuttle, B.R.; Bording, J.K.; Kodambaka, S. Mechanical stability of possible structures of PtN investigated using first-principles calculations. Phys. Rev. B 2006, 73. [Google Scholar] [CrossRef]
  58. Born, M.; Huang, K. Dynamical Theory of Crystal Lattices; Oxford University Press: London, UK, 1954. [Google Scholar]
  59. Wu, Z.J.; Zhao, E.J.; Xiang, H.P.; Hao, X.F.; Liu, X.J.; Meng, J. Crystal structures and elastic properties of superhard IrN2 and IrN3 from first principles. Phys. Rev. B 2007, 76. [Google Scholar] [CrossRef]
  60. Zhang, B.; Wu, L.L.; Wan, B.; Zhang, J.W.; Li, Z.H.; Gou, H.Y. Structural evolution, mechanical properties, and electronic structure of Al-Mg-Si compounds from first principles. J. Mater. Sci. 2015, 50, 6498–6509. [Google Scholar] [CrossRef]
  61. Zhou, W.; Liu, L.J.; Li, B.L.; Wu, P.; Song, Q.G. Structural, elastic and electronic properties of intermetallics in the Pt-Sn system: A density functional investigation. Comput. Mater. Sci. 2009, 46, 921–931. [Google Scholar] [CrossRef]
  62. Chung, D.H.; Buessem, W.R. The elastic anisotropy of crystals. J. Appl. Phys. 1967, 38, 2010–2012. [Google Scholar] [CrossRef]
  63. Chung, D.H.; Buessem, W.R. The Voigt-Reuss-Hill approximation and elastic moduli of polycrystalline MgO, CaF2, β-ZnS, ZnSe, and CdTe. J. Appl. Phys. 1967, 38, 2535–2540. [Google Scholar] [CrossRef]
  64. Chung, D.H.; Buessem, W.R. The Voigt-Reuss-Hill (VRH) approximation and the elastic moduli of polycrystalline ZnO, TiO2 (rutile), and α-Al2O3. J. Appl. Phys. 1968, 39, 2777–2782. [Google Scholar] [CrossRef]
  65. Xiang, H.P.; Wu, Z.J. Ab initio study on the electronic, magnetic, and mechanical properties of CaCu3V4O12. Inorg. Chem. 2008, 47, 2706–2709. [Google Scholar] [CrossRef] [PubMed]
  66. Chen, X.Q.; Niu, H.Y.; Li, D.Z.; Li, Y.Y. Modeling hardness of polycrystalline materials and bulk metallic glasses. Intermetallics 2011, 19, 1275–1281. [Google Scholar] [CrossRef]
  67. Li, Y.F.; Gao, Y.M.; Xiao, B.; Min, T.; Fan, Z.J.; Ma, S.Q.; Xu, L.L. Theoretical study on the stability, elasticity, hardness and electronic structures of W-C binary compounds. J. Alloys Compd. 2010, 502, 28–37. [Google Scholar] [CrossRef]
  68. Teter, D.M. Computional alchemy: The search for new superhard materials. MRS Bull. 1998, 23, 22–27. [Google Scholar] [CrossRef]
  69. Chen, K.; Zhao, L. Elastic properties, thermal expansion coefficients and electronic structures of Ti0.75X0.25C carbides. J. Phys. Chem. Solids 2007, 68, 1805–1811. [Google Scholar] [CrossRef]
  70. Šimůnek, A.; Vackář, J. Hardness of covalent and ionic crystals: First-principle calculations. Phys. Rev. Lett. 2006, 96. [Google Scholar] [CrossRef] [PubMed]
  71. Pabst, W.; Tichá, G.; Gregorová, E. Effective elastic properties of alumina-zirconia composite ceramics part 3. Calculation of elastic moduli of polycrystalline alumina and zirconia from monocrystal data. J. Ceram.-Silikáty 2004, 48, 41–48. [Google Scholar]
  72. Birch, F. Finite elastic strain of cubic crystals. Phys. Rev. 1947, 71, 809–824. [Google Scholar] [CrossRef]
  73. Manghnani, M.H.; Fisher, E.S.; Brower, W.S., Jr. Temperature dependence of the elastic constants of single-crystal rutile. J. Phys. Chem. Solids 1972, 33, 2149–2159. [Google Scholar] [CrossRef]
  74. Otero-de-da-Roza, A.; Luaña, V. GIBBS2: A new version of the quasi-harmonic model code. I. Robust treatment of the static data. Comput. Phys. Commun. 2011, 182, 1708–1720. [Google Scholar] [CrossRef]
  75. Nye, J.F. Physical Properties of Crystals; Oxford University Press: London, UK, 1985. [Google Scholar]
  76. Ravindran, P.; Fast, L.; Korzhavyi, P.A.; Johansson, B.; Wills, J.; Eriksson, O. Density functional theory for calculation of elastic properties of orthorhombic crystals: Application to TiSi2. J. Appl. Phys. 1998, 84, 4891–4904. [Google Scholar] [CrossRef]
  77. Xiang, H.M.; Feng, Z.H.; Zhou, Y.C. Theoretical investigations on mechanical anisotropy and intrinsic thermal conductivity of YbAlO3. J. Eur. Ceram. Soc. 2015, 35, 1549–1557. [Google Scholar] [CrossRef]
  78. Ranganathan, S.I.; Ostoja-Starzewski, M. Universal elastic anisotropy index. Phys. Rev. Lett. 2008, 101. [Google Scholar] [CrossRef] [PubMed]
  79. Mersol, S.A.; Lynch, C.T.; Vahldiek, F.W. Anisotropy in Single Crystal Refractory Compounds; Plenum Press: New York, NY, USA, 1968. [Google Scholar]
  80. Brugger, K. Determination of third-order elastic coefficients in crystals. J. Appl. Phys. 1965, 36, 768–773. [Google Scholar] [CrossRef]
  81. Grimvall, G. Thermophysical Properties of Materials; Elsevier: Amsterdam, The Netherlands, 1999. [Google Scholar]
  82. Feng, J.; Xiao, B.; Wan, C.L.; Qu, Z.X.; Huang, Z.C.; Chen, J.C.; Zhou, R.; Pan, W. Electronic structure, mechanical properties and thermal conductivity of Ln2Zr2O7 (Ln = La, Pr, Nd, Sm, Eu and Gd). Acta. Mater. 2011, 59, 1742–1760. [Google Scholar] [CrossRef]
  83. Chong, X.Y.; Jiang, Y.H.; Zhou, R.; Zhu, H.; Feng, J. Electronic structure, anisotropic elastic and thermal properties of the η phase Fe6W6C. Comput. Mater. Sci. 2015, 108, 205–211. [Google Scholar] [CrossRef]
  84. Ding, Y.C.; Xiao, B. Anisotropic elasticity, sound velocity and thermal conductivity of TiO2 polymorphs from first principles calculations. Comput. Mater. Sci. 2014, 82, 202–218. [Google Scholar] [CrossRef]
  85. Anderson, O.L. A simplified method for calculating the Debye temperature from elastic constants. J. Phys. Chem. Solids 1963, 24, 909–917. [Google Scholar] [CrossRef]
  86. Kittel, C. Introduction to Solid State Physics, 8th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2004. [Google Scholar]
  87. Rocha, F.S.D.; Fraga, G.L.F.; Brandao, D.E.; Silva, C.M.D.; Gomes, A.A. Specific heat and electronic structure of Heusler compounds Ni2TAl (T = Ti, Zr, Hf, V, Hb, Ta). Physica B 1999, 269, 154–162. [Google Scholar] [CrossRef]
  88. Clarke, D.R. Materials selection guidelines for low thermal conductivity thermal barrier coatings. Surf. Coat. Technol. 2003, 163, 67–74. [Google Scholar] [CrossRef]
  89. Cahill, D.G.; Watson, S.K.; Pohl, R.O. Lower limit to the thermal conductivity of disordered crystals. Phys. Rev. B 1992, 46, 6131–6140. [Google Scholar] [CrossRef]
  90. Zheng, K.H.; Gao, Y.M.; Li, Y.F.; Zhao, S.M.; Wang, J. Three-body abrasive wear resistance of iron matrix composites reinforced with ceramic particles. Proc. Inst. Mech. Eng. Part J J. Eng. Tribol. 2014, 228, 3–10. [Google Scholar] [CrossRef]
  91. ASM International Handbook Committee. ASM Handbook, Vol. 1: Properties and Selection: Irons, Steels and High-Performance Alloys; ASM International: Geauga County, OH, USA, 1993. [Google Scholar]
  92. Onderka, B.; Sypien, A.; Wierzbicka-Miernik, A.; Czeppe, T.; Zabdyr, L.A. Specific heat capacities of some ternary aluminides. J. Phase Equilib. Diffus. 2011, 32, 39–41. [Google Scholar] [CrossRef]
  93. Drulis, M.K.; Czopnik, A.; Drulis, H.; Spanier, J.E.; Ganguly, A.; Barsoum, M.W. On the heat capacity of Ti3GeC2. Mater. Sci. Eng. B 2005, 119, 159–163. [Google Scholar] [CrossRef]
Figure 1. Crystal structure calculated for AlNi2Ti compounds.
Figure 1. Crystal structure calculated for AlNi2Ti compounds.
Crystals 08 00093 g001
Figure 2. Temperature dependence of the lattice parameter and bulk modulus of AlNi2Ti.
Figure 2. Temperature dependence of the lattice parameter and bulk modulus of AlNi2Ti.
Crystals 08 00093 g002
Figure 3. Compressibility characterized by normalized volume versus pressure of AlNi2Ti, as well as diamond, α-Al2O3, BCC-iron, TiO2 and TiO for comparison.
Figure 3. Compressibility characterized by normalized volume versus pressure of AlNi2Ti, as well as diamond, α-Al2O3, BCC-iron, TiO2 and TiO for comparison.
Crystals 08 00093 g003
Figure 4. The surface constructions of the bulk modulus B and Young’s modulus E of AlNi2Ti; the corresponding planar projections of B on the [001] crystallographic plane and planar projections of E on [001] and [110] crystallographic planes are shown below.
Figure 4. The surface constructions of the bulk modulus B and Young’s modulus E of AlNi2Ti; the corresponding planar projections of B on the [001] crystallographic plane and planar projections of E on [001] and [110] crystallographic planes are shown below.
Crystals 08 00093 g004
Figure 5. Dependence of the Helmholtz free energy E(V, T) on the crystal volume of AlNi2Ti under different temperatures.
Figure 5. Dependence of the Helmholtz free energy E(V, T) on the crystal volume of AlNi2Ti under different temperatures.
Crystals 08 00093 g005
Figure 6. Volume expansions evaluated by V/V0 as a function of temperature, where V0 refers to the equilibrium volume at 0 K.
Figure 6. Volume expansions evaluated by V/V0 as a function of temperature, where V0 refers to the equilibrium volume at 0 K.
Crystals 08 00093 g006
Figure 7. Linear thermal expansion coefficients of AlNi2Ti and Ni-Ti binary compounds as a function of temperature [52].
Figure 7. Linear thermal expansion coefficients of AlNi2Ti and Ni-Ti binary compounds as a function of temperature [52].
Crystals 08 00093 g007
Figure 8. The specific heats at constant pressure (Cp) or volume (Cp) of AlNi2Ti as a function of temperature [92].
Figure 8. The specific heats at constant pressure (Cp) or volume (Cp) of AlNi2Ti as a function of temperature [92].
Crystals 08 00093 g008
Figure 9. The heat capacity of AlNi2Ti and Ni-Ti binary compounds plotted in the range of 0–45 K.
Figure 9. The heat capacity of AlNi2Ti and Ni-Ti binary compounds plotted in the range of 0–45 K.
Crystals 08 00093 g009
Figure 10. Total and partial density of states of AlNi2Ti compound. The Fermi level is indicated by a red vertical line.
Figure 10. Total and partial density of states of AlNi2Ti compound. The Fermi level is indicated by a red vertical line.
Crystals 08 00093 g010
Table 1. The optimized ground-state structural parameters (a, b, c, in Å; Vcell in Å3; ρ in g/cm3), atomic site, cohesive energy (Ecoh, eV/f.u.) and formation enthalpy (ΔrH, eV/atom) of AlNi2Ti intermetallic compound.
Table 1. The optimized ground-state structural parameters (a, b, c, in Å; Vcell in Å3; ρ in g/cm3), atomic site, cohesive energy (Ecoh, eV/f.u.) and formation enthalpy (ΔrH, eV/atom) of AlNi2Ti intermetallic compound.
MethodLattice ParametersVcellρAtomic Site Δ r H Ecoh
a = b = cAlNiTi
GGA-PBE5.908 (5.865 a)206.242 (201.75 a)6.193 (6.33 a)4a (0, 0, 0)8c (0.25, 0.25, 0.25)4b (0.5, 0.5, 0.5)−0.671−27.602
GGA-PW915.903205.6766.210−0.643−33.173
Other cal. data5.906 b, 5.90 c, 5.828 d −0.658 b, −0.662 m
Exp. data5.906 e, 5.883–5.910 f, 5.819 g, 5.876 i, 5.942 j, 5.889 k, 5.865–5.886 l203.61–206.40 f −0.747 e, −0.578 h, −0.579 n
EOS5.911206.4996.186
a From Inorganic Crystal Structure Database (ICSD) #58063; b Calculation data evaluated by GGA in [35]; c Cal. data evaluated by Reddy, et al. in [36]; d Cal. data evaluated by Monte Carlo simulations in [11]; e Exp. data in [47]; f Exp. data from [48]; g Data calculated from the selected area electron diffraction pattern in [22]; h Exp. data from [49]; i Data from handbook [50]; j Exp. data measured by X-ray diffraction in [51]; k Exp. data in [52]; l Exp. data in [53]; m Calculated values from a thermodynamic database in [54]; n Data measured by high-temperature reaction calorimeter in [54].
Table 2. Calculated independent elastic constants (GPa) compared to experimental and other theoretical values for the single crystalline AlNi2Ti intermetallic compound. Hereafter the mechanical modulus (GPa), Poisson’s ratio, Vicker’s hardness, shear anisotropic factors, and percent anisotropic index are also provided.
Table 2. Calculated independent elastic constants (GPa) compared to experimental and other theoretical values for the single crystalline AlNi2Ti intermetallic compound. Hereafter the mechanical modulus (GPa), Poisson’s ratio, Vicker’s hardness, shear anisotropic factors, and percent anisotropic index are also provided.
MethodElastic ConstantsBGBH/GHEvHVShear Anisotropic FactorsAnisotropic Index
c11c12c44BVBRBHGVGRGHA1A2A3ABAGAU
GGA-PBE208.1142.297.6164.2164.2164.271.854.763.22.6168.10.324.42.962.962.960.000.141.56
GGA-PW91209.3143.494.8165.4165.4165.470.154.262.12.7165.60.334.12.882.882.880.000.131.47
Cal. data a215.8 a, 223 b139.4 a, 135 b98.7 a, 104 b164.9 a, 164 b67.5 a, 74 b2.4 a178.1 a, 193 b 8.1 a2.36 b 0.104 a1.165 a
Exp. data120 c97 c56 c
EOS 162.9
a Cal. data from [35]; b Cal. data from [36]; c Data adjusted by comparing with experimental diffraction elastic constants in [18];
Table 3. Calculated anisotropic sound velocities of AlNi2Ti intermetallic compounds, together with the values of Ni3Ti. The unit of velocity is m/s.
Table 3. Calculated anisotropic sound velocities of AlNi2Ti intermetallic compounds, together with the values of Ni3Ti. The unit of velocity is m/s.
Species[100] [010][001][110][111]νlνtνm
νlνt1νt2νlνt1νt2νlνt1νt2νlνt1νt2νlνt1νt2
AlNi2Ti579739713971579739713971579739713971663732623971689429672967633431953583
Ni3Ti656532773625656532773625656532773277656532773625724311723513642735503955
Table 4. Calculated thermodynamic parameters: Debye temperatures (K), and minimum thermal conductivities (k, W/(m·K)) of the AlNi2Ti intermetallic compound and Ni3Ti for comparison.
Table 4. Calculated thermodynamic parameters: Debye temperatures (K), and minimum thermal conductivities (k, W/(m·K)) of the AlNi2Ti intermetallic compound and Ni3Ti for comparison.
SpeciesModelΘD M ¯ (10−26)P (1028)[100] kmin[010] kmin[001] kmin[110] kmin[111] kminkmink (Exp.)
AlNi2TiClark455.1 (462 a, 411 b)7.97 0.850.850.851.161.371.1421.4 c
Cahill 7.761.391.391.391.401.301.29
Ni3TiClark516.69.29 1.361.361.451.361.261.31
Cahill 8.431.441.441.401.441.281.45
a Cal. data from [35]; b Exp. data from [87]; c Exp. data from [2].
Table 5. Theoretically calculated low-temperature specific heat of AlNi2Ti and Ni-Ti binary compounds, including the total density of states at the Fermi level (Df, states/(eV·atom)), the characteristic parameters of electron (γ, mJ/(K2·mol)) and phonon (β, mJ/(K4·mol)) specific heat, and the specific heat (Cp, J/(K·mol)) at 3 K and 30 K for comparison.
Table 5. Theoretically calculated low-temperature specific heat of AlNi2Ti and Ni-Ti binary compounds, including the total density of states at the Fermi level (Df, states/(eV·atom)), the characteristic parameters of electron (γ, mJ/(K2·mol)) and phonon (β, mJ/(K4·mol)) specific heat, and the specific heat (Cp, J/(K·mol)) at 3 K and 30 K for comparison.
SpeciesDfγβCp (3 K)Cp (30 K)
AlNi2Ti0.4160.9810.08250.005172.26
Ni3Ti0.1130.2670.05690.002341.54
B2_NiTi0.7141.680.05770.006611.61
B19’_NiTi0.4060.9570.08060.005052.21
NiTi20.7631.800.1150.008503.16

Share and Cite

MDPI and ACS Style

Tang, S.; Li, Y.; Gao, Y.; Zheng, Q.; Liu, Z.; Ren, X. First-Principles Investigations of the Structural, Anisotropic Mechanical, Thermodynamic and Electronic Properties of the AlNi2Ti Compound. Crystals 2018, 8, 93. https://doi.org/10.3390/cryst8020093

AMA Style

Tang S, Li Y, Gao Y, Zheng Q, Liu Z, Ren X. First-Principles Investigations of the Structural, Anisotropic Mechanical, Thermodynamic and Electronic Properties of the AlNi2Ti Compound. Crystals. 2018; 8(2):93. https://doi.org/10.3390/cryst8020093

Chicago/Turabian Style

Tang, Shuli, Yefei Li, Yimin Gao, Qiaoling Zheng, Zhiwei Liu, and Xiangyi Ren. 2018. "First-Principles Investigations of the Structural, Anisotropic Mechanical, Thermodynamic and Electronic Properties of the AlNi2Ti Compound" Crystals 8, no. 2: 93. https://doi.org/10.3390/cryst8020093

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop