† Corresponding author. E-mail:
First-principles computation on the basis of density functional theory (DFT) is executed with the CASTEP code to explore the structural, elastic, and electronic properties along with Debye temperature and theoretical Vickers’ hardness of newly discovered ordered MAX phase carbide Mo2TiAlC2. The computed structural parameters are very reasonable compared with the experimental results. The mechanical stability is verified by using the computed elastic constants. The brittleness of the compound is indicated by both the Poisson’s and Pugh’s ratios. The new MAX phase is capable of resisting the pressure and tension and also has the clear directional bonding between atoms. The compound shows significant elastic anisotropy. The Debye temperature estimated from elastic moduli (B, G) is found to be 413.6 K. The electronic structure indicates that the bonding nature of Mo2TiAlC2 is a mixture of covalent and metallic with few ionic characters. The electron charge density map shows a strong directional Mo–C–Mo covalent bonding associated with a relatively weak Ti–C bond. The calculated Fermi surface is due to the low-dispersive Mo 4d-like bands, which makes the compound a conductive one. The hardness of the compound is also evaluated and a high value of 9.01 GPa is an indication of its strong covalent bonding.
The MAX phases are a class of compounds that have a common formula Mn+1AXn with n = 1, 2, 3, etc. Here M denotes an early transition metal from group 3–6, A is an A-group element that comes from columns 12–16 in the periodic table, and X represents C and/or N. The MAX phases are the materials with laminated layered structures having thickness values of the individual layers on a nanometer scale. This family of ternary layered carbides and nitrides are promising materials carrying the merits of both metals and high-performance ceramics.[1] These metallic ceramics were first discovered in the 1960s by Nowotny et al.[2–8] All the studies completed over the past six decades have confirmed that these nanolaminates exhibit an uncommon arrangement of properties, which put together the high specific stiffness and lubricity with assuagement of machinability and good oxidation resistance including expedient high-temperature properties. In addition, the MAX phases possess surprisingly good electrical and thermal conductivities, exceptional damage tolerance, excellent thermal shock resistance, and reversible plasticity, high resistance to oxidation and corrosion and ability to maintain the strengths to high temperature.[9–15] This rare collection of metallic and ceramics properties of the MAX compounds makes them aspirant materials for a series of applications ranging from catalysis to aerospace.[16]
At present, we have more than 70 synthesized MAX phases amongst which there is only the Mo2GaC phase containing Mo as M element. Very recently, Anasori et al.[17] reported on the synthesis of a new Mo-containing ordered MAX phase, Mo2TiAlC2 having structure similar to Ti3SiC2.[10,18] In this phase, the Ti atoms are pressed between two Mo-layers like a layer pressed between two others of a different kind in a sandwich form that in turn are contiguous to the Al planes resulting in an Mo–Ti–Mo–Al–Mo–Ti–Mo stacking order. The C-atoms take positions in between the Mo and Ti layers. In the present study, the first-principles method is used to predict the elastic and electronic properties along with Debye temperature and theoretical hardness of this new member of the MAX family. Very recently, we became aware of a report on various properties of newly discovered Mo2TiAlC2 with first-principles calculations by the projector augmented wave method as implemented within the Vienna ab-initio simulation package (VASP).[19] The present results are compared with those found in this report.
The structural, elastic, and electronic properties of the newly synthesized MAX phase Mo2TiAlC2 are calculated by using the plane wave pseudopotential approach within the density functional theory (DFT) that is used in the Cambridge Serial Total Energy Package (CASTEP) code.[20] The generalized gradient approximation (GGA) using the Perdew–Burke–Ernzerhof (PBE) functional[21] is used to evaluate the exchange–correlation energy. The Vanderbilt-type ultrasoft pseudopotential is taken for the treatment of electron–ion interactions.[22] The energy cut-off is set to be 550 eV to expand the plane wave functions. The first Brillouin zone of the unit cell is sampled by using the Monkhorst–Pack scheme[23] of k-points with 17 × 17 × 2 mesh. The structure is fully optimized with respect to atomic positions and lattice parameters by means of the Broyden–Fletcher–Goldfarb–Shanno (BFGS) algorithm.[24] The tolerances of convergence for the energy, maximum force, maximum stress, and maximum atomic displacement are set to be 5 × 10−6 eV/atom, 0.01 eV/Å, 0.02 GPa, and 5 × 10−4 Å, respectively. To achieve good convergences in total energy, geometry, and elastic moduli, the parameters set above cover the required adequacy. To obtain the smooth Fermi surface, 30 × 30 × 4 k-point mesh is used.
The structure of Mo2TiAlC2 is first optimized. The elastic constants are calculated from the first-principles via the finite-strain theory as implemented in the CASTEP code. In the finite-strain theory, a set of finite identical deformations is applied and the resulting stress is computed to find the elastic constants by solving the equation, σi = Cijεj. The detailed procedures can be found in Ref. [25]. The polycrystalline elastic properties namely, bulk modulus B and shear modulus G are calculated from the single-crystal elastic constants Cij using the well-known Voigt–Reuss–Hill approximation.[26–28] In addition, the Young’s modulus Y, Poisson’s ratio v, and shear anisotropy factor A are derived by using the equations Y = (9GB)/(3B + G), v = (3B − 2G)/(6B + 2G), and A = 4C44/(C11 + C33 − 2C13), respectively.
The Debye temperature, θD, is calculated by using one of the standard methods, depends on the elastic constants such as bulk modulus and shear modulus.[29] Within this method, the Debye temperature can be estimated from the average sound velocity vm by the following equation:
The theoretical Vickers’ hardness for crystal with metallic bonding is evaluated from the empirical formula:[30,31]
Mo2TiAlC2 crystallizes in a hexagonal structure with a space group of P63/mmc and is isostructural with Ti3SiC2. Figure
The elastic properties are directly related to the crystal structure and the nature of bonding between atoms within the system. These in turn largely determine the phonon spectrum and the Debye temperature of the compound. The predicted single crystal elastic constants Cij under 0 GPa are presented in Table
The polycrystalline elastic properties such as bulk modulus B, shear modulus G, Young’s modulus Y, Pugh’s ratio G/B, Poisson’s ratio v, and shear elastic anisotropy factor A derived from Cij are also listed in Table
Elastic anisotropy of a crystal reflects a characteristic feature of bonding in different directions. Inherently, most of the known crystals are elastically anisotropic, and a precise depiction of such an anisotropic character has, therefore, an essential implication in crystal physics and science of engineering. It shows a relationship with the possibility of existence of microcracks in the crystals. For hexagonal crystals, a shear anisotropy factor associated with the (100) shear plane between the 〈011〉 and 〈010〉 directions is defined as A = 4C44/(C11 + C33 − 2C13). For an isotropic crystal, A is found to be unity. The deviation of A from unity assesses the elastic anisotropy possessed by the crystal and the amount of deviation measures the level of elastic anisotropy. The shear anisotropic factor for Mo2TiAlC2 is given in Table
The Debye temperature θD is an influential parameter directly related to various physical properties including melting temperature and specific heat, and is used to make a distinction between high- and low-temperature regions for a solid material. The Debye temperature well determines a demarcation between quantum and classical behavior of phonons. When the temperature T of a solid is raised over θD, all modes of vibrations are expected to have energy equal to kBT. At T < θD, the high-frequency modes are found to be stationary. The vibrational excitations at low temperature appear only from acoustic vibrations. Consequently, the Debye temperature calculated from elastic constants is seen to be the same as that estimated from measured specific heat.[38] The Debye temperature along with the related quantities including longitudinal, transverse and average sound velocities calculated within the present formalism is listed in Table
A proper depiction of electronic structure is essential to explain many physical phenomena such as optical spectra of materials. The calculated electronic energy band structure of Mo2TiAlC2 at equilibrium lattice parameters along the high-symmetry lines in the first Brillouin zone is shown in Fig.
To give a clear view of the nature of the electronic band structure, the total and partial energy density of states (DOS) are also calculated and explained as indicated in Fig.
The charge transfer and Mulliken atomic populations are also calculated in order to obtain the bonding characters in Mo2TiAlC2. Listed in Table
Therefore, it can be seen that the Mo–C bond is more covalent than the Ti–C bond, and Ti–C bond holds stronger covalent bonding than the Mo–Al bond. The values of degree of metallicity[31,40] defined as fm = Pμ′/Pμ for Mo–C, Ti–C, and Mo–Al bonds are 0.021, 0.033, and 0.063, respectively. Hence, the metallicity ranking of the bonds is Mo–Al > Ti–C > Mo–C; the Mo–Al bond has the highest metallicity, yielding strong metallic bonding. Thus, based on the above discussion, one can determine that the bonding nature of the new compound is dominated by the covalent and metallic bonds. The calculated theoretical hardness is also listed in Table
To further clarify the nature of chemical bonding in Mo2TiAlC2, the electron charge density distribution is investigated and the contour of electron charge density in
The contour of electron charge density reveals a strong directional Mo–C–Mo covalent bond chain with each pair of the chains linked by a relatively weak Ti–C bond. Because of a large difference in electronegativity, the electronic charge around Mo atoms is attracted towards C atoms and a strong covalent–ionic bonding along Mo and C direction is induced. The hybridized Mo 4d–C 2p states are, in fact, responsible for the forming of these bonds. Additionally, the electron charge density of Mo just overlaps with that of Al, which is an indication of a relatively weak bonding between Mo and Al. The present results are in good agreement with the findings that MAX phases characteristically have the remarkably strong M–X bonds and rather weak M–A bonds.[1]
Figure
The structural, elastic, and electronic properties of Mo2TiAlC2 are studied by the DFT-based first-principles pseudopotential total energy method. The calculated structural parameters agree fairly with both the experimental and theoretical data. The single crystal elastic constants ensure the mechanical stability of the new MAX nanolaminate by satisfying the Born criteria. The Poisson’s and Pugh’s ratios suggest that Mo2TiAlC2 should behave as a brittle material. The new ordered MAX carbide has the ability to resist the pressure and tension. The compound is characterized by significant elastic anisotropy. The shear modulus and Debye temperature as well as bond overlap population and Vickers’ hardness indicate the strong directional bonding between atoms in the compound. The chemical bonding is seen to be a combination of covalent, metallic and ionic nature. The electron charge density map reveals that the electronic charge around Mo atoms is attracted towards C atoms and a strong covalent–ionic bonding along Mo and C direction is observed The investigated Fermi surface originates mainly from the low-dispersive Mo 4d-like bands, which is responsible for the conductivity of Mo2TiAlC2. Finally, we hope that these theoretical results inspire experimental research to measure the elastic properties, Debye temperature and Vickers’ hardness of the newly discovered layered MAX phase.
1 | |
2 | |
3 | |
4 | |
5 | |
6 | |
7 | |
8 | |
9 | |
10 | |
11 | |
12 | |
13 | |
14 | |
15 | |
16 | |
17 | |
18 | |
19 | |
20 | |
21 | |
22 | |
23 | |
24 | |
25 | |
26 | |
27 | |
28 | |
29 | |
30 | |
31 | |
32 | |
33 | |
34 | |
35 | |
36 | |
37 | |
38 | |
39 | |
40 | |
41 |