Band structure, Fermi surface, elastic, thermodynamic, and optical properties of AlZr3, AlCu3, and AlCu2Zr: First-principles study
Parvin R1, Parvin F1, Ali M S2, †, , Islam A K M A3
Department of Physics, Rajshahi University, Rajshahi-6205, Bangladesh
Department of Physics, Pabna University of Science and Technology, Pabna 6600, Bangladesh
International Islamic University Chittagong, 154/A College Road, Chittagong-4203, Bangladesh

 

† Corresponding author. E-mail: shahajan199@yahoo.com

Abstract
Abstract

The electronic properties (Fermi surface, band structure, and density of states (DOS)) of Al-based alloys AlM3 (M = Zr and Cu) and AlCu2Zr are investigated using the first-principles pseudopotential plane wave method within the generalized gradient approximation (GGA). The structural parameters and elastic constants are evaluated and compared with other available data. Also, the pressure dependences of mechanical properties of the compounds are studied. The temperature dependence of adiabatic bulk modulus, Debye temperature, specific heat, thermal expansion coefficient, entropy, and internal energy are all obtained for the first time through quasi-harmonic Debye model with phononic effects for T = 0 K–100 K. The parameters of optical properties (dielectric functions, refractive index, extinction coefficient, absorption spectrum, conductivity, energy-loss spectrum, and reflectivity) of the compounds are calculated and discussed for the first time. The reflectivities of the materials are quite high in the IR–visible–UV region up to ∼ 15 eV, showing that they promise to be good coating materials to avoid solar heating. Some of the properties are also compared with those of the Al-based Ni3Al compound.

1. Introduction

Aluminum-based alloys have received much attention in recent years because of their important roles in many technological applications. They have vast applications in the aerospace, microelectronic, motorized vehicles, and domestic industry. As the alloys are important in the engineering and research areas, mechanical behavior has become a focus of various investigations. A large quantity of alloy elements can be used to form the Al-based intermetallic compounds. It is observed that the bonding behaviors and electronic natures are changed of such a type of alloys which show excellent mechanical, physical, and chemical properties.[1] The compounds involving aluminum and transition metals are known to have high resistance to oxidation and corrosion, elevated-temperature strength, relatively low density, and high melting points, which make them desirable candidates for high-temperature structural applications.[2] Zirconium can effectively enhance the mechanical strengths of the alloys when copper and zinc elements exist in aluminum and Al-based alloys.[3] Many investigations have focused on the constituent binary systems, such as Al–Cu, Al–Zr, and Cu–Zr.[49] Some theoretical investigations have been performed on Al-based alloys such as structural, elastic and electronic properties.[10,11]

To the best of our knowledge, electronic band structures, Fermi surfaces, thermal and optical properties of AlZr3, AlCu3, and AlCu2Zr compounds have not yet been discussed theoretically and experimentally. The thermodynamic properties include a variety of properties and phenomena such as bulk modulus, specific heat, thermal expansion coefficient, Debye temperature, etc. The specific heat of a material is one of the most important thermodynamic properties indicating its heat retention or loss ability. We have also briefly discussed our calculated structural and elastic properties for AlZr3, AlCu3, and AlCu2Zr intermetallic alloys.

On the other hand, the optical properties of solids provide an important tool for studying energy band structures, impurity levels, localized defects, lattice vibrations, and certain magnetic excitations. The optical conductivity or the dielectric function indicates a response of a system of electrons to an applied field. Thus there is a need to deal with all these issues which have been covered for the first time in the present paper. A detailed discussion and analysis are made, some of which are in comparison with the results of available Al-based alloys.[10,12]

2. Computational details

The density functional theory (DFT) based on the CASTEP code has been used in the present investigation.[1316] The interactions between valence electrons and ions are considered using the ultrasoft pseudo-potential formalism of Vanderbilt.[17] To find out the minimum potential energy of the compound, a quasi-Newton (variable-metric) minimization method, the Broyden–Fletcher–Goldfarb–Shanno (BFGS) update scheme,[18] is utilized, which provides a very efficient and fast way to explore the optimizing crystal structure. The electron wave function is expanded in plane waves up to an energy cutoff of 400 eV for AlZr3 and 500 eV for AlCu3 as well as AlCu2Zr for the sampling of the Brillouin zone,[19] a 15×15×15, 21×21×21, 9×9×9 Monkhorst–Pack mesh are employed[20] for AlZr3, AlCu3, and AlCu2Zr, respectively. Generalized gradient approximation (GGA) of the Perdew–Burke–Ernzerhof for solids (PBEsol) formalism,[16] which is dependent on both the electron density and its gradient at each space point is used to evaluate the exchange-correlation energy. Optimization is performed using a convergence threshold of 10−5 eV/atom for the total energy and 10−3 Å for maximum displacement. Maximum force and stress are 0.03 eV/Å and 0.05 GPa, respectively for all the calculations. Norm conserving pseudo-potential is used only for calculating the optical properties, keeping the other setup unchanged. The resulting stress can be calculated with respect to the optimized crystal structure[21] using a set of homogeneous deformations with a finite value.

The thermodynamic properties are studied within the quasi-harmonic Debye model implemented in the Gibbs program,[22] a detailed description of which can be found in Refs. [23]– [25]. We use the DFT calculated EV (energy ∼ volume) data at T = 0 K, P = 0 GPa, and the third-order Birch–Murnaghan EOS.[26] Using this, the thermodynamic properties such as the bulk modulus, thermal expansion co-efficient, specific heats, and Debye temperature at non-zero temperature and pressure are calculated.

3. Results and discussion
3.1. Structural properties

Aluminum-based intermetallic AlM3 (M = Zr and Cu) binary alloys crystallize into the cubic Cu3Au structure[27,28] with space group Pm-3m (No. 221) consisting of Al atoms at the corners and M atoms at the face centers of the cube. The atomic positions are Al: 1a (0,0,0), and M: 3c (0,1/2,1/2). The ternary AlCu2Zr alloy is a partially ordered Cu2MnAl-type fcc (space group No. 225: Fm-3m) structure,[29] where Al atoms occupy the 4a Wyckoff site (0,0,0), Cu atoms occupy the 8c site (1/4,1/4,1/4) and Zr atoms occupy the 4b site (1/2,1/2,1/2). In order to investigate the ground-state properties of AlM3 (M = Zr and Cu) and AlCu2Zr alloys, the geometry optimizations are performed first with the full relaxation of the cell shape and atomic positions. The equilibrium crystal structures of these compounds are shown in Fig. 1. The results of first-principles calculations of the structural properties of these compounds are presented in Table 1, together with the available experimental values[2729] and other theoretical results.[10] The comparison shows that our results are in reasonable agreement with both theoretical and experimental values.

Fig. 1. Crystal structures of AlM3 (M = Zr and Cu) and AlCu2Zr.
Table 1.

Values of calculated and experimental lattice constant a (in unit Å), equilibrium volume V0 (in units Å3), bulk modulus B0 (in unit GPa) of AlZr3, AlCu3, and AlCu2Zr alloys.

.
3.2. Mechanical properties

Generally, elastic constants of a solid permit us to obtain the mechanical properties and they can be used to describe the resistance of a crystal to an externally applied stress. They also provide important information about bonding characteristics near the equilibrium state. Thus, it is essential to investigate the elastic constants to understand the mechanical behavior of AlM3 (M = Zr and Cu) and AlCu2Zr compounds at equilibrium and under pressure. For a material with cubic symmetry, there are only three independent elastic constants, i.e., C11, C12, and C44. In this paper, the calculated single crystal elastic constants of these compounds are listed in Table 2. It can be seen that the obtained elastic constants of the alloys agree well with the theoretical results in Ref. [10], which indicates that the DFT method we used in this study is reasonable. Furthermore, the corresponding requirements of mechanical stability in a cubic crystal lead to the following restriction on the elastic constants:[30] C11 +2C12 > 0, C44 > 0, C11C12 > 0. As shown in Table 2, all of the elastic constants for the compounds satisfy these criteria mentioned above, indicating that these compounds are mechanically stable.

Table 2.

Values of calculated elastic constants Cij (in unit GPa), bulk modulus B (in unit GPa), shear modulus G (in unit GPa), Young’s modulus E (in unit GPa), Poisson’s ratio ν, elastic anisotropic factor A for AlZr3, AlCu3, and AlCu2Zr alloys at equilibrium.

.

The pressure dependences of the calculated elastic constants of AlZr3, AlCu3, and AlCu2Zr are plotted in Fig. 2. The linear increases of elastic constants are due to the decreases of lattice parameters of all the compounds under pressure (not shown). It is noticeable that C11 increases more rapidly with pressure than C12 and C44. C11 represents elasticity in the x direction, i.e., in length, the longitudinal strain produces a significant change in volume. Again, the volume change is highly related to the pressure and thus produces a large change in C11. Conversely, the constants C12 and C44 are related to the elasticity in shape.

Fig. 2. Pressure dependences of elastic constants of (a) AlZr3, (b) AlCu3, and (c) AlCu2Zr.

The polycrystalline elastic properties, such as bulk modulus B, shear modulus G, Young’s modulus E, and Poisson’s ratio ν are estimated from the calculated single crystal elastic constants, and are presented in Table 2. The calculated results show that G for AlZr3 is the largest, while for AlCu2Zr it is less than that for AlCu3, which agrees well with the available theoretical calculation.[10]

The pressure effects on the calculated bulk modulus, shear modulus, and Young’s modulus of AlZr3, AlCu3, and AlCu2Zr are shown in Fig. 3(a). The rate of increase of G with pressure is smaller than those of B and E. The shear modulus G is mainly due to the elastic constant C44, because a large C44 implies a stronger resistance to shear in the (100) plane. Young’s moduli of AlZr3 and AlCu2Zr are nearly equal at P = 0 GPa but when pressure increases the Young’s modulus for AlCu2Zr increases more rapidly than those for the other two phases, which indicates that AlCu2Zr phase is much stiffer than the other two phases at high pressure.

Fig. 3. Pressure effects on (a) bulk modulus (B), shear modulus (G), and Young’s modulus (E); (b) the values of ratio of bulk modulus to shear modulus (B/G) and Poisson’s ratio (ν) of AlZr3, AlCu3, and AlCu2Zr.

Figure 3(b) shows the pressure dependences of Poisson’s ratio and B/G of AlZr3, AlCu3, and AlCu2Zr. Poisson’s ratio can reflect the bonding properties of material. For the different bonding materials, Poisson’s ratios are different.[31] As for a covalent material, the value of ν is small (typically ∼ 0.1); for an ionic material, the typical value of ν is 0.25; for a metallic material, ν is typically 0.33.[31] In this case, Poisson’s ratio ranges from 0.264 to 0.341, which indicates that the AlM3 (M = Zr and Cu) and AlCu2Zr compounds show metallic behaviors. Poisson’s ratio increases with pressure increasing, the larger Poisson’s ratio signifies better plasticity. Hence it can be said that AlCu3 possesses better plasticity than the other two phases. Additionally, in order to evaluate the ductility of material, a parameter B/G has been proposed by Pugh.[32] The critical value separating ductility from brittleness is about 1.75. In the present work, the calculated values of B/G (Table 2) for these three compounds are all higher than 1.75, implying that AlM3 (M = Zr and Cu) and AlCu2Zr alloys behave in a ductile manner. Figure 3(b) shows that B/G increases with pressure growing for all the compounds under study. It indicates that AlZr3, AlCu3, and AlCu2Zr are ductile compounds at up to 50 GPa. Anisotropy is used to describe situations where properties vary systematically, dependent on the direction. A = 2C44/(C11C12) is called a shear anisotropy factor, often used to represent the elastic anisotropy of crystals.[33] The A values of 1.52, 1.89, and 1.30 for AlM3 (M = Zr and Cu) and AlCu2Zr alloys, respectively quantify elastic anisotropies. It is noted that AlCu3 phase is much more anisotropic than the other two phases.

3.3. Electronic band structure and Fermi surface

The investigated electronic band structures along high symmetry directions in the Brillouin zones together with the total densities of states (DOSs) of AlZr3, AlCu3, and AlCu2Zr are depicted in Fig. 4. The band structures reveal metallic characters with dispersion bands crossing the Fermi level (EF) for all structures. For AlZr3 and AlCu2Zr alloys, Zr 4d orbitals are mainly responsible for the metallic characters. But for AlCu3, at the Fermi level, the main contribution comes from the Cu 3d orbitals.

Fig. 4. Band structures and total DOSs of (a) AlZr3, (b) AlCu3, and (c) AlCu2Zr.

It is observed from Fig. 4(a) that for AlZr3 alloy, the hybridization between Al 3p orbital and Zr 4d orbital is strong both below and above the Fermi level. For this reason the entire DOS can be divided into bonding and anti-bonding regions and therefore a pseudogap resides in between. The characteristic pseudogap around the Fermi level indicates the presence of the directional covalent bonding. The Fermi level located at a valley in the bonding region implies that the system has a pronounced stability.[10] There is a general consideration that the formation of covalent bonding would enhance the strength of material in comparison with the pure metallic bonding.[34]

According to the covalent approach, the guiding principle is to maximize bonding. Hence the stability increases with increasing occupancy in the bonding region, for a series of compounds having the same structure.[35] Therefore it can be said that AlZr3 is structurally more stable than AlCu3. On the other hand, for AlCu2Zr, the main bonding peaks from Fermi level to 3 eV predominantly derive from the Zr 4d orbitals. The phase stability of intermetallic compound depends on the location of the Fermi level and the value of the DOS at EF.[36,37] The lower value of DOS at EF corresponds to a more stable structure. The values of the total DOS at the Fermi level in states are 3.2 eV and 9.3 eV for AlZr3 and AlCu2Zr alloys, respectively. Therefore, it can be said that AlZr3 has a more stable structure than the other two alloys of Al under consideration, which agrees very well with the observation in Ref. [10].

The shape of the Fermi surface is derived from the periodicity and symmetry of the crystalline lattice and from the occupation of electronic energy bands. It separates unfilled orbitals from the filled orbitals. The Fermi surface topologies of AlZr3, AlCu3, and AlCu2Zr are shown in Fig. 5. The shapes of the Fermi surfaces are quite different for the three Al-based intermetallic compounds. In the cases of AlZr3 and AlCu3, two energy bands cross the Fermi level. They form two Fermi sheets. AlZr3 has a hole-like Fermi surface around R point as well as around X point. The topology of the Fermi surface of AlZr3 is nearly similar to that of YPb3.[38] Again, AlCu3 has two hole-like Fermi sheets around X point. On the other hand, AlCu2Zr alloy has four Fermi sheets because four energy bands cross the Fermi level. The topology of this Fermi surface is complicated; it has hole-like sheets around R point and electron-like sheets around Γ point.

Fig. 5. Fermi surfaces of (a) AlZr3, (b) AlCu3, and (c) AlCu2Zr.

Charge transport is mainly limited by carrier–carrier and carrier–phonon scattering. Besides, there is always impurity scattering in most real systems. The carrier–carrier scattering is intimately related to the electronic band structure and density of states at the Fermi level. The topology of the Fermi surface also plays a role. For example, the highly dispersive nature of energy dispersion curve for AlZr3 (Fig. 4(a)) around the Fermi energy, implies high carrier mobility. This is supported by the large low energy optical conductivity seen in Fig. 11(h) for the same compound (AlZr3).

3.4. Thermal properties

The thermodynamic properties such as the bulk moduli, specific heats, Debye temperatures, and volume thermal expansion coefficients at different temperatures and pressures are evaluated for the first time for these compounds. We calculate all the thermodynamic properties in a temperature range from 0 K to 1000 K and a pressure range from 0 GPa to 50 GPa, where the quasi-harmonic Debye model remains valid.

In the quasi-harmonic model, the nonequilibrium Gibbs function G∗(V;P,T) can be written as[22]

where E(V) is the total energy per unit cell, PV corresponds to the constant hydrostatic pressure condition, and Avib(ΘD(V);T) is the vibrational Helmholtz free energy which is given by

where n is the number of atoms per formula unit, k is the Boltzmann constant, D(ΘD/T) is the Debye integral and ΘD represents the Debye temperature. The thermal equation of state V(P,T) and the chemical potential G(P,T) can be obtained by minimizing the nonequilibrium Gibbs function with respect to volume V. Other macroscopic properties as a function of pressure P and temperature T can also be derived from standard thermodynamic relations.[22]

The temperature dependences of adiabatic bulk modulus, B of AlZr3, AlCu3, and AlCu2Zr alloys at P = 0 GPa are shown in Fig. 6(a). Our calculations exhibit that B value for AlCu2Zr is nearly flat whereas for AlZr3 compound, it decreases with temperature increasing up to 1000 K. We also note that in a temperature range of 0 K–600 K, B fluctuates for AlCu3, it then decreases smoothly with temperature increasing until 1000 K. For AlZr3, AlCu3, and AlCu2Zr alloys, the values of B are found respectively to be 99 GPa, 126 GPa, and 129 GPa at P = 0 GPa and T = 0 K, which are in good agreement with the values computed using elastic constant data (Table 2).

Fig. 6. Temperature dependences of (a) adiabatic bulk modulus and (b) Debye temperature of AlZr3, AlCu3, and AlCu2Zr.

Debye temperature is basically a measure of the vibrational response of the material. Debye temperature is not a strictly determined parameter; various estimates may be obtained through well-established empirical or semi-empirical formula, relating ΘD to various macroscopic properties.[39] Figure 6(b) depicts the temperature dependences of Debye temperature, ΘD of AlZr3, AlCu3, and AlCu2Zr compounds at zero pressure. Our calculated values of ΘD using the quasi-harmonic Debye model are 367 K, 377 K, and 476 K, respectively for AlZr3, AlCu3, and AlCu2Zr alloys at zero pressure and zero temperature, which are in reasonable agreement with the values of 385 K, 381 K, and 397 K estimated using elastic constants data. These may be compared with the calculated value of ΘD for Ni3Al is 466 K.[12]

It is observed from Fig. 6(b) that ΘD is the highest for AlCu2Zr and the lowest for AlZr3 at zero temperature and pressure. The Debye temperatures decrease slowly with increasing temperature for AlCu2Zr and AlZr3. But for AlCu3, the Debye temperature decreases faster than those of the other two compounds. The temperature variations of all the phases reveal that there is a strong bonding in both the AlCu2Zr and AlZr3 compounds where a weak bonding occurs in AlCu3, which coincides with the DOS and elastic properties. This is due to the fact that the Debye temperature is related to the volume, V and adiabatic bulk modulus, B. The adiabatic bulk modulus B decreases and increases with increasing temperature and pressure, respectively but the volume, V, behaves in an opposite manner. It is found in Ref. [22] that B makes more contribution to the Debye temperature than, V. On the other hand, Debye temperature increases with increasing pressure (not shown).

The specific heat of the material is usually approximated by the contribution of the lattice specific heat. Figure 7 presents the estimated results on the temperature dependences of constant-pressure and constant-volume specific heat capacities CP, CV of AlZr3, AlCu3, and AlCu2Zr alloys. Due to the increase of temperature, phonon thermal softening occurs and hence heat capacities increase. The differences between CP and CV for AlZr3, AlCu3, and AlCu2Zr alloys are due to the thermal expansion caused by anharmonicity effects.[40] In the low temperature limit, the specific heat exhibits the Debye T3 power-law behavior; at high temperature (T > 300 K), the anharmonic effect on heat capacity is suppressed, and CV approaches the classical asymptotic limit CV = 3nNkB = 99.7 J/mol·K. These results exhibit the fact that the interactions between ions in AlZr3, AlCu3, and AlCu2Zr alloys have large effects on heat capacities especially at low temperatures. The calculations of Fatmi et al.[12] for a similar type of compound coincide with our results. To evaluate the electronic contribution to specific heat through the Sommerfeld constant, γ within the free electron model: , we can use N(EF) from the investigated DOS for the three alloys. This scheme gives the values 7.55, 2.95, and 21.93 mJ·mol−1·K−2 for AlZr3, AlCu3, and AlCu2Zr alloys, respectively.

Fig. 7. Specific heats at constant (a) pressure and (b) volume versus temperature of AlZr3, AlCu3, and AlCu2Zr.

Finally, the computed results on the variations of the volume thermal expansion coefficient, αv with temperature and pressure for AlZr3, AlCu3, and AlCu2Zr alloys are displayed in Fig. 8. The volume thermal expansion coefficient, αv reflects the temperature dependence of volume at constant pressure: α = (1/V)(dV/dT)P. Figure 8(a) depicts the temperature dependence of αv at P = 0 GPa. The coefficient, αv increases rapidly as the temperature increases up to ∼ 300 K and then gradually approaches towards a linear increase with temperature rising and the propensity of increment becomes moderate, which means that the temperature dependence of αv is very small at high temperature. For a given temperature, the coefficient decreases drastically with the increase of pressure. The pressure dependence of αv at 300 K is presented in Fig. 8(b). At high temperatures and high pressures, the thermal expansion would converge to a constant value. With temperature rising, the average amplitude of atomic vibrations increases. For an anharmonic potential, it corresponds to the increase in the average value of inter-atomic separation, i.e., thermal expansion. The stronger the inter-atomic bonding, the smaller the thermal expansion is. The values of αv are found to be 7.5×10−6 K−1, 3.1×10−5 K−1, and 5.3×10−5 K−1 for AlCu2Zr, AlZr3, and AlCu3 alloys, respectively at P = 0 GPa and T = 300 K. The higher value of AlCu3 indicates that it is softer than the other two compounds. For a given temperature, the coefficient αv sharply decreases with the increase of pressure but for AlCu2Zr, it is almost constant (see Fig. 8(b)). To date, there have been no experimental results nor other theoretical investigations for the thermal properties of AlZr3, AlCu3, and AlCu2Zr alloys; our investigations may serve as a guide for future research.

Fig. 8. Variations of αv with (a) temperature and (b) pressure of AlZr3, AlCu3, and AlCu2Zr.

The variations of entropy, S of AlZr3, AlCu3, and AlCu2Zr alloys each as a function of temperature and pressure are shown in Fig. 9. It is observed from Fig. 9(a) that at P = 0 GPa, S increases with increasing temperature. The rates of increase of entropy with temperature are almost the same for the compounds under study. This type of diagram is used in thermodynamics because it helps to visualize the heat transfer in the process. Entropy is a measure of how much the energies of atoms and molecules become more spread out in a process. Figure 9(b) shows that the entropy decreases with increasing pressure. This is due to the fact that when the pressure on the system increases, the volume decreases. The energies of the particles are in a smaller space, so they are less spread out. Hence the entropy decreases. It is observed from Fig. 9(b) that the rate of decrease of S is the largest for AlCu3 and the smallest for AlCu2Zr. Therefore the volume of AlCu3 is more compressible than those of the other two Al-based alloys under the same pressure.

Fig. 9. Variations of S with (a) temperature and (b) pressure for AlZr3, AlCu3, and AlCu2Zr.

The temperature and pressure variations of the calculated internal energies for AlZr3, AlCu3, and AlCu2Zr alloys are shown in Fig. 10. The internal energy is the energy associated with the random, disordered motions of molecules. It is observed from Figs. 10(a) and 10(b) that the internal energies increase with temperature and pressure increasing. At P = 0 GPa the rates of increase of U with temperature increasing are the same for all the compounds under study. On the other hand, at room temperature (T = 300 K), the internal energies increase 5.62%, 6.68%, and 4.72% for AlZr3, AlCu3, and AlCu2Zr alloys respectively, when the pressure changes from 0 GPa to 50 GPa.

Fig. 10. (a) Temperature and (b) pressure dependences of U for AlZr3, AlCu3, and AlCu2Zr.
3.5. Optical properties

The optical properties of AlZr3, AlCu3, and AlCu2Zr alloys may be derived from the knowledge of the complex dielectric function, ɛ(ω). Generally, there are two contributions for ɛ(ω), namely from intraband and interband transitions. The contribution from the latter is vital only for metals.[41] The complex dielectric function can be expressed as ɛ(ω) = ɛ1(ω)+ iɛ2(ω). The imaginary part ɛ2(ω) is obtained from the momentum matrix elements between the occupied and unoccupied wave functions within the selection rules, whereas the real part ɛ1(ω) of dielectric function can be derived from the imaginary part using the Kramers–Kronig relation. The knowledge of both ɛ1(ω) and ɛ2(ω) allows the calculation of important optical constants such as the refractive index n(ω), extinction coefficient k(ω), conductivity, optical reflectivity R(ω), absorption coefficient α(ω), and energy-loss spectrum L(ω) using the expressions given in Ref. [42]. The calculational approaches are well established and widely available in the literature[43,44] and hence will not be repeated here.

The estimated optical properties of AlZr3, AlCu3, and AlCu2Zr alloys at ground state are presented in Fig. 11, for the photon energy ranging from 0 eV up to 20 eV for polarization vector [100]. The available theoretical data for Ni3Al[12] are also shown in the plots for comparison. Dielectric function is the most general property of a solid, which modifies the incident electromagnetic wave of light. Figures 11(a) and 11(b) show the real and imaginary parts of the dielectric function for AlZr3, AlCu3, and AlCu2Zr alloys along with the calculated values of Ni3Al. Our results are similar to those for Ni3Al.[12] The electronic band structure analysis shows that AlZr3, AlCu3, and AlCu2Zr alloys are metallic. Therefore, Drude term correction is required to include the effect of metallicity.[45,46]

Fig. 11. (a) Real and (b) imaginary parts of dielectric function, (c) real and (d) imaginary parts of refractive index, (e) absorption, (f) loss function, (g) reflectivity, and (h) real part of conductivity of AlZr3, AlCu3, and AlCu2Zr.

The plasma frequencies of 10 eV, 4 eV, 6 eV, and damping values of 0.07 eV, 0.05 eV, 0.02 eV are used for AlZr3, AlCu3, and AlCu2Zr alloys, respectively. For all calculations, we use a 0.5-eV Gaussian smearing. The real part ɛ1 goes through zero from below about 14.52 eV, 21.66 eV, 12.47 eV and the imaginary part ɛ2 approaches zero damping values of 14.98 eV, 22.80 eV, and 12.98 eV for AlZr3, AlCu3, and AlCu2Zr alloys, respectively. Metallic reflectance characteristics are exhibited in the range ɛ1 < 0.

The velocity of propagation of an electromagnetic wave through a solid is given by the frequency dependent complex refractive index N = n + ik, where the real part n is related to the velocity and the imaginary part k, the extinction coefficient is related to the decay or damping of the oscillation amplitude of the incident electric field. The refractive indexes and extinction coefficients are illustrated in Figs. 11(c) and 11(d). The static refractive index n(0) is found to have the values 11.8 (AlCu3), 19.3 (AlZr3), and 12.4 (AlCu2Zr), which agrees well with that of another Al-based intermetallic compound Ni3Al (14.56).[12] The values of extinction coefficients, k each show a maximum value at 0.0 eV, then they decrease rapidly and turn to zero at about 15.0 eV, 22.96 eV, and 12.43 eV for AlZr3, AlCu3, and AlCu2Zr alloys, respectively.

Figure 11(e) shows the absorption coefficient spectra of the AlZr3, AlCu3, and AlCu2Zr alloys, which begin at zero photon energy due to their metallic nature. In Fig. 11(e), we show the absorption coefficient spectrum with one peak between 0.78 eV and 5.73 eV for AlZr3 which rises and then decreases rapidly in the high-energy region. Nearly the same features can be seen for AlCu3 and AlCu2Zr but with two prominent peaks for AlCu3. The highest peaks appear at 7.55 eV, 6.26 eV, and 5.80 eV for AlZr3, AlCu3, and AlCu2Zr alloys, respectively. The absorption coefficient spectrum of Ni3Al is also shown in Fig. 11(e) for comparison, for which the highest peak occurs at 14.93 eV.[12] It is observed from the figure that at higher photon energies, Ni3Al and AlCu3 have roughly similar absorption coefficient spectra.

Loss function is intimately related to absorption and reflection. Loss function refers to the fast electron traversing in a material. It is clear from the study of the figure that maximum reflectivity occurs where the minimum absorption takes place. So the study of loss function is the most important in material study. The electron energy loss function L(ω) is depicted in Fig. 11(f). The peaks in L(ω) spectra represent the characteristics associated with the plasma resonance and the corresponding frequency is the so-called plasma frequency, ωP,[47] which occurs at ɛ2 < 1 and ɛ1 = 0.[46,48] The peaks of L(ω) located at 14.5 eV, 21.7 eV, and 12.4 eV correspond to plasma frequencies of AlZr3, AlCu3, and AlCu2Zr alloys, respectively. Further, these peaks correspond to irregular edges in the reflectivity spectra (Fig. 11(g)), and hence an abrupt reduction occurs at each of these peak values in the reflectivity spectra. If the incident light has frequency greater than the plasma frequencies of the alloys under study, these alloys will be transparent and will change from metallic to dielectric response.

The variations of calculated optical reflectivity R(ω) with incident photon energy are displayed in Fig. 11(g). It is observed from the figure that AlZr3 and AlCu2Zr have their maximum reflectivity values (∼100%) in the infra-red region as well as in the ultra-violet region. On the other hand, AlCu3 has maximum reflectivity only in the infra-red region. Again in the visible light region (energy range ∼1.8 eV–3.1 eV), it is observed from Fig. 11(g) that AlCu3 has larger reflectivity than those of the AlZr3 and AlCu2Zr alloys, but all of them have reflectivities more than 44%. This indicates that the capability of AlCu3 to reflect visible light is stronger than those of the other two alloys. On the other hand, AlZr3 and AlCu2Zr alloys are better reflectors than that of the AlCu3 alloy in the ultra-violet region. According to Li et al.,[45] a MAX-phase compound will be capable of reducing solar heating if it has a reflectivity of ∼ 44% in the visible light region. Hence, we may conclude that all the compounds under study are also candidate materials for coating to reduce solar heating.

For the three alloys, their photoconductivities start with zero photon energy (Fig. 11(h)). This indicates that the materials have no band gap that is evident from band structure calculations (Fig. 2). Moreover, the photoconductivities and hence electrical conductivities of these materials increase as a result of absorbing photons. From Fig. 11(h), we observe that the maximum optical conductivities occur at energy 2.72 eV for AlZr3, 5.40 eV for AlCu3, and 3.21 eV for AlCu2Zr within the energy range of 2 eV to 6.84 eV, respectively. This implies that AlCu3 will be more highly electrically conductive than the other two alloys when the incident radiation has an energy within this range. The photoconductivities reach zero at about 10 eV for both the compounds AlZr3 and AlCu2Zr; AlCu3 also reaches zero but at 22.5 eV (not shown). There is a photoconductivity for none of the phases when the photon energy is higher than 22.5 eV. We are not aware of any other theoretical calculations or experimental data for the optical properties of AlZr3, AlCu3, and AlCu2Zr alloys. Thus, our results would provide a useful reference for further investigations.

4. Conclusions

First-principles calculations based on DFT and the quasi-harmonic Debye model are used to investigate different properties of AlZr3, AlCu3, and AlCu2Zr alloys. Based on the calculated results, we find that AlZr3 phase has the highest stiffness among the three phases. On the other hand, the Poisson’s ratio changes from 0.264 to 0.341, which indicates that AlM3 (M = Zr and Cu) and AlCu2Zr compounds show metallic behaviors. In view of Pugh’s criterion for these three alloys it is observed that these three compounds show ductility behavior. Under the effect of pressure, the value of C11 increases rapidly with pressure compared to C12 and C44 for all compounds, which indicates that the longitudinal strain produces a significant change in volume. The larger Poisson’s ratio of AlCu3 compared to those of the other two phases indicates better plasticity.

It is also observed from the DOS value that AlZr3 has a more stable structure than the other two alloys. The pseudogap around the Fermi level indicates the presence of the directional covalent bonding. The AlCu3 and AlZr3 show hole-like Fermi surface whereas AlCu2Zr shows both electron-like and hole-like Fermi surface. The stronger the inter-atomic bonding, the smaller the thermal expansion is. The smaller value of AlCu2Zr indicates stronger atomic bonding whereas the higher value of AlCu3 indicates that it is softer than the other two compounds. The reflectivities of these compounds are high (∼100%) in the IR as well as in the UV region. In the visible region, all the compounds under study may be considered as potential coating materials to avoid solar heating.

Reference
1Sauthoff GWestbrook J HFleischer R L1994Intermetallic CompoundsNew YorkWiley 1 991
2Cahn R W1998Intermetallics6563
3Rajagopalan P KSharma I GKrishnan T S1999J. Alloys Compd.285212
4Zhou WLiu L JLi B LSong Q GWu P2009J. Electron. Mater.38356
5Emmanuel CSanchez J M2002Phys. Rev. B65094105
6Ghosh GAsta M2005Acta Mater.533225
7Ghosh G2007Acta Mater.553347
8Ma W JWang Y RWei B CSun Y F2007Trans. Nonferrous Met. Soc. China17929
9Pauly SDas JMattern NKim D HEckert J2009Intermetallics17453
10Guan Y ZZhang H YLi W2011Physica B4061149
11Zhang B RJia ZDuan X ZYang X Z2013Acta Phys. Polonica A23668
12Fatmi MGhebouli M AGhebouli BChihi TBoucetta SHeiba Z K2011Rom. J. Phys.56935
13Clark S JSegall M DProbert M JPickard C JHasnip P JPayne M C2005Z. Kristallogr.220567
14Materials Studio CASTEP manual 2010 ©Accelrys 2010 p. 261 http://www.tcm.phy.cam.ac.uk/castep/documentation/WebHelp/CASTEP.html
15Hohenberg PKohn W1964Phys. Rev.136864
16Perdew J PRuzsinszky ACsonka G IVydrov O AScuseria G EConstantin L AZhou XBurke K2008Phys. Rev. Lett.100136406
17Vanderbilt D1990Phys. Rev. B417892
18Head J DZerner M C1985Chem. Phys. Lett.122264
19Brillouin C RHebd C R1930Seances. Acad. Sci.191292
20Monkhorst H JPack J D1976Phys. Rev. B135188
21Milman VWarren M C2001J. Phys.: Condens. Matter13241
22Blanco M AFrancisco ELuaña V2004Comput. Phys. Commun.15857
23Blanco M APenda’s A MFrancisco ERecio J MFranco RMol J1996Struct. Theochem.368245
24Flórez MRecio J MFrancisco EBlanco M APendás A M2002Phys. Rev. B66144112
25Francisco ERecio J MBlanco M APendás A M1998J. Phys. Chem.1021595
26Birch F1978J. Geophys. Res.831257
27Meng W JFaber JJrOkamoto P RRehn L EKestel B JHitterman R L1990J. Appl. Phys.671312
28Draissia MDebili M YBoukhris NZadam MLallouche S2007Copper1065
29Meyer Z UReckendorf RSchmidt P CWeiss A Z1989Phys. Chem. N F163103
30Mattesini MAhuja RJohansson B2003Phys. Rev. B68184108
31Haines JLeger J MBocquillon G2001Ann. Rev. Mater. Res.311
32Pugh S F1954Philos. Mag.45823
33Zener C M1948Elasticity and Anelasticity of MetalsChicagoUniversity of Chicago Press
34Chen PLi D LYi J XLi WTang B YPeng L M2009Solid State Sci.112156
35Xu J HLin WFreeman A J1993Phys. Rev. B484276
36Xu J HOguchi TFreeman A J1987Phys. Rev. B364186
37Hong TWatson-Yang T JFreeman A JOguchi TXu J H1990Phys. Rev. B4112462
38Ram SwetarekhaKanchana VSvane ADugdale S BChristensen N E2013J. Phys.: Condens. Matter25155501
39Leadbetter H MReep R PClark A F1983OhioASM Metal Park
40Ali M SIslam A K M AHossain M MParvin F2012Physica B4074221
41Reshak A HAtuchin V VAuluck SKityk I V2008J. Phys.: Condens. Matter20325234
42Shaha SSinha T PMookarjee A2000Phys. Rev. B628828
43Li CWang BLi YWang R2009J. Phys. D: Appl. Phys.42065407
44Li CKuo JWang BLi YWang R2009J. Phys. D: Appl. Phys.42075404
45Li SAhuja RBarsoum M WJena PJohansson B2008Appl. Phys. Lett.92221907
46Saniz RYe L HShishidou TFreeman A J2006Phys. Rev. B74014209
47Fox M1972Optical Properties of SolidsNew YorkAcademic Press
48De Almeida J SAhuja R2006Phys. Rev. B73165102