Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

First-principles insights into thermo-mechanical, electronic, and optical properties of Hf2AC (A = Cl, Br) MAX phases through A-site halogen tuning

Asif Mohammed Arfiae, Mahamudun Nabib, Md. Hasan Mia*c, Omar Alsalmid, Muhammad Athar Uddine and Md. Zahid Hasan*f
aDept of Computer Science and Engineering, University of Yamanashi, Takeda, 4-4-37 Kofu, Yamanashi, Japan
bDepartment of Physics, University of Rajshahi, Rajshahi 6205, Bangladesh
cDepartment of Computer and Communication Engineering, International Islamic University Chittagong Kumira, Chattogram, 4318, Bangladesh. E-mail: mdhasan111.ru@gmail.com
dDepartment of Physics, College of Science, Umm Al-Qura University, Makkah 21955, Saudi Arabia
eDepartment of Electrical and Electronic Engineering, International Islamic University Chittagong, Kumira, Chittagong, 4318, Bangladesh
fMaterials Research and Simulation Lab, Department of Electrical and Electronic Engineering, International Islamic University Chittagong, Kumira, Chittagong, 4318, Bangladesh. E-mail: zahidhasan.02@gmail.com

Received 25th September 2025 , Accepted 27th October 2025

First published on 5th November 2025


Abstract

In this study, we conduct a comprehensive first-principles investigation of the mechanical, electronic, optical, and thermo-mechanical properties of Hf2AC MAX phases, where the A-site halogen is varied between Cl and Br. We use density functional theory to look into how changing the A-site atom affects the structural stability and other functional properties of these layered ternary carbides. Structural analysis reveals that replacing Cl with the larger Br atom increases lattice constants and unit cell volume, aligning with the atomic size trend. Both phases satisfy key stability criteria, including thermodynamic, mechanical, and dynamic stability, confirmed by formation energies, elastic constants, and phonon dispersions. Mechanical property analysis shows that Hf2BrC has slightly reduced stiffness compared to Hf2ClC due to weaker Hf–Br bonding, though both maintain ductility and mechanical robustness. Thermo-mechanical parameters, such as Debye temperature, melting point, and thermal conductivity, are influenced by halogen mass and bonding character. Thermo-lattice properties, including heat capacity and the Grüneisen parameter, further support their thermal stability over a broad temperature range. The electronic band structures and density of states show that both compounds act like metals, with Hf d-orbitals at the Fermi level being the most important. Optical properties, derived from the dielectric function, indicate strong activity in the visible and ultraviolet regions. Br substitution causes a red shift in absorption and reflectivity spectra, enhancing plasmonic and photonic behavior. High reflectivity and photoconductivity beyond the plasma frequency highlight their potential in optoelectronic and UV shielding applications. The ability to fine-tune these materials through atomic-level modifications opens new pathways for designing MAX phases with tailored performance for use in coatings, electronics, and energy-related applications.


1. Introduction

MAX phases, a unique class of layered ternary carbides and nitrides, have garnered significant scientific interest owing to their remarkable combination of metallic and ceramic characteristics. These materials are structurally resilient at high temperatures, exhibit excellent machinability, and demonstrate superior electrical and thermal conductivity along with exceptional resistance to thermal shock, corrosion, and oxidation.1–3 Such a blend of properties makes MAX phases ideal for a wide array of high-performance industrial applications, including machinable refractories, heating elements, rotating electrical contacts, bearings, nozzles, die-pressing tools, and heat exchangers. Notably, their ability to withstand neutron irradiation has also led to investigations into their potential roles in nuclear reactor components.4 In recent years, several MAX phases have also been explored as potential candidates for thermal barrier coatings (TBCs), given their favorable thermal and mechanical behavior at elevated temperatures.5,6

Structurally, MAX phases are characterized by a hexagonal layered lattice and are described by the general chemical formula Mn+1AXn, where M is an early transition metal, A is an element from groups 13 to 17 of the periodic table, and X is either carbon or nitrogen.2 The integer n indicates the number of M–X layers in the structure; for n = 1, the resulting formula becomes M2AX, known as the 211-type MAX phase, which is the focus of this investigation.2,5,7

Over 60 stable ternary MAX phases have been experimentally synthesized, and numerous solid solutions have been reported through substitution at the M and A sites. Common examples include (Nb, Zr)2AlC, (Ti, V)2SC, (Ti, Hf)2SC, V2(Al, Ga)C, Ti2(Si, Ge)C, and Cr2(Al, Ge)C.8–15 However, A-site solid solutions are typically limited to elements from groups 13 to 16, such as Al, Si, or Ga. Recently, a new class of MAX phases has emerged where halogen elements (Cl, Br, I) replace the conventional A-site elements. These include compounds such as La2Br(C, N), La2Cl(C, N), Sc2ClC, Sc2BrC, Hf2ClC, and Hf2BrC, which indicate the structural feasibility of incorporating halogens at the A site in MAX-like frameworks.2

The unique characteristics of M2AX phases, commonly referred to as “211” MAX phases, have captivated the scientific community due to their remarkable combination of metallic and ceramic properties. These materials exhibit a fascinating array of attributes: moderate hardness (Vickers hardness of 2.0–8.0 GPa), anisotropy, elastic stiffness, excellent machinability, damage tolerance, and high-temperature plasticity.16 They are lightweight, resistant to fatigue, creep, oxidation, and thermal shock, and demonstrate a transition from brittleness to plasticity at elevated temperatures. These extraordinary features make MAX phases highly desirable for diverse technological applications. Recent studies have further expanded the understanding of M2AX phases, showcasing their multifunctionality. For instance, H. Kim et al.17 identified M2AX phases (M = lanthanides, A = Al or Si, X = C or N) as exceptional candidates for bond coat applications. Similarly, O. S. Rijal et al.18 explored the properties of Ti2CdC and Ti2SC, revealing their potential in optoelectronics, advanced photonic devices, solar thermal collectors, and thermal insulation. The works of B. U. Haq et al.19 and M. N. Amin et al.20 have highlighted the applications of W2GaX and Cr2AC phases, respectively, in high-temperature environments, optoelectronic devices, and other advanced technologies. Furthermore, W. Tan et al.21 demonstrated the potential of S-based M2SX phases in thermoelectric applications, while M. A. Ali et al.22 and M. M. Uddin et al.23 underscored the suitability of Hf2SeC, Hf2AlC, and Hf2AlN for high-temperature and solar heating reduction applications.

Despite these advances, a comprehensive understanding of how halogen substitution at the A site influences the physical properties of MAX phases is still lacking.

The prior work by Ohmer et al.2 focused on predicting the elastic constants and moduli of Hf2ClC and Hf2BrC within a broad high-throughput screening of 82 potential MAX phases, using basic thermodynamic descriptors to assess phase stability. While this study was foundational in identifying the existence of these compounds, it primarily addressed phase feasibility rather than functional behavior. In contrast, our present study provides a comprehensive, in-depth, and comparative investigation of the functional properties influenced by halogen substitution in Hf-based MAX phases, extending well beyond prior structural predictions and providing helpful information regarding their mechanical, electronic, thermal, and optical behavior. We systematically explore their thermo-mechanical responses by analyzing elastic anisotropy indices, Debye and melting temperatures, the Grüneisen parameter, and both lattice and minimum thermal conductivities, which are critical for understanding their thermal resilience and structural integrity. Additionally, we discuss their electronic structure through detailed assessments of band dispersion, directional charge carrier mobility, effective mass, and the nature of orbital hybridization—key insights necessary for electronic and energy-related applications. Furthermore, we present a complete evaluation of the optical behavior of these materials, including their dielectric functions, absorption characteristics, photoconductivity, reflectivity, and energy-loss spectra, which have not been previously reported. To the best of our knowledge, no earlier study has comprehensively examined the collective influence of halogen substitution (Cl versus Br) on such an extensive range of physical properties in Hf-based MAX phases. Our findings address this critical gap and present a valuable framework for the property-driven design of advanced MAX materials, highlighting their potential across multifunctional domains such as thermal barrier coatings, optoelectronics, and structural applications operating under extreme environments.

The arrangement of this paper is as follows: Section 2 describes the DFT computational setting. Section 3 presents and analyzes observations and their physical implications. Section 4 provides a brief summary of the present findings and concluding remarks.

2. DFT computational setting

In this work, we explore the impact of halogen substitution, specifically Cl and Br, at the A site of Hf2AC (A = Cl, Br) MAX phases through comprehensive first-principles calculations based on density functional theory (DFT).24 All calculations were performed using the CASTEP code,25 a plane-wave-based density functional theory (DFT) software package. Ultrasoft pseudopotentials were employed for all atomic species (Hf, Cl, Br, and C), allowing accurate treatment of valence electron interactions while maintaining computational efficiency. To accurately capture the exchange–correlation interactions, we employed the generalized gradient approximation (GGA) using the Perdew–Burke–Ernzerhof (PBE) functional.26 A plane-wave energy cutoff of 500 eV was selected to ensure convergence of total energy across all investigated systems. For Brillouin zone integration, a Monkhorst–Pack k-point mesh of 9 × 9 × 2 was used, offering an optimal balance between computational efficiency and numerical accuracy.27 The interactions between valence electrons and ionic cores were described using ultrasoft pseudopotentials, which were carefully chosen to ensure reliable modeling of chemical bonding.28 The valence states considered in our calculations were Hf (5d25s2), Cl (3s23p5), Br (4s24p5), and C (2s22p2), which play a critical role in defining the material properties of the Hf2ClC and Hf2BrC compounds. Geometry optimizations were carried out using the Broyden–Fletcher–Goldfarb–Shanno (BFGS) minimization algorithm.29 The convergence thresholds were rigorously set to a total energy difference of 5 × 10−6 eV per atom, a maximum force of 0.01 eV Å−1, a maximum stress of 0.02 GPa, and a maximum displacement of 5 × 10−4 Å. These strict criteria ensured that the optimized structures were robust and well-suited for subsequent property evaluations. Following structural optimization, we systematically computed various mechanical properties of Hf2ClC and Hf2BrC, including their elastic constants via the stress–strain method.30 Phonon dynamic properties, including the phonon dispersion and phonon density of states (PDOS), were calculated using the finite displacement method.31 The crystallographic structures were visualized using the VESTA software package, providing insights into the atomic arrangements and symmetry.32

3. Observations and physical implications

3.1 Crystallographic properties

The substitution of halogen (Cl, Br) in Hf2AC (A = Cl, Br) MAX phases induce significant changes in both the structural characteristics and properties of these compounds due to differences in atomic size, electronegativity, and bonding behavior. As illustrated in Fig. 1, these compounds crystallize in the hexagonal crystal system with the space group P63/mmc (no. 194), with each unit cell containing eight atoms and two formula units. The atomic positions for Hf2AC (A = Cl, Br) are defined by the Wyckoff coordinates: C atoms occupy the 2a site at (0, 0, 0), A-site atoms (Cl or Br) are located at 2c (1/3, 2/3, 1/4), and Hf atoms are positioned at 4f (2/3, 1/3, ZM). The crystal structure is determined by two main lattice parameters, a and c, along with the internal atomic coordinate ZM.
image file: d5ra07272g-f1.tif
Fig. 1 Optimized crystal structures of Hf2AC (A = Cl, Br) MAX phases: (a) standard orientation highlighting the layered hexagonal architecture, (b) projection along the a-axis emphasizing the stacking sequence, and (c) projection along the c-axis revealing the atomic arrangement within the basal plane.

The primary effect of substituting Cl and Br in the A-site on Hf2AC MAX phases is observed in the c/a ratio and unit cell volume. Hf2BrC exhibits a slightly higher c/a ratio and unit cell volume than Hf2ClC, which is consistent with the larger atomic radius of Br compared to Cl. The trend of increasing c/a ratio and unit cell volume with larger atoms is further confirmed by comparing Hf2ClC and Hf2BrC with other Hf-based MAX phases in Table 1. These changes in the crystal structure reflect a typical trend observed in MAX phases, where the incorporation of larger atoms leads to expanded interlayer spacing and a corresponding increase in unit cell volume. The equilibrium crystal structures of Hf2AC (A = Cl, Br) have been optimized by minimizing the total energy, with the resulting structural parameters provided in Table 1. Although direct experimental or theoretical benchmarks for Hf2ClC and Hf2BrC are currently unavailable, our optimized values show strong consistency with those reported for other well-known and experimentally verified MAX phases, thereby reinforcing the reliability of our calculations and their alignment with established trends in the field (see Table 1).

Table 1 Structural and energetic properties of Hf2AC (A = Cl, Br) MAX phases, featuring optimized lattice parameters (a, c in Å), c/a ratio, internal lattice parameter (ZM in Å), unit cell volume (V in Å3), formation enthalpy (ΔHf in eV), formation energy (ΔEf in eV) and coordination sites (Or, Pr) along with theoretical comparisons
Phases a c c/a ZM V ΔHf ΔEf Or Pr Ref.
Hf2ClC 3.432 12.947 3.77 0.0834 132.094 −7.13 −1.26 1.77 1.17 This
Hf2BrC 3.490 13.427 3.84 0.0833 141.636 −6.95 −1.10 1.75 1.16 This
Hf2AlC 3.275 14.362 4.39 0.0876           Expt.36
Hf2SeB 3.522 12.478 3.54             Expt.37
Hf2AlC 3.310 14.350 4.33 0.0890 136.345   Theo.23
Hf2SC 3.423 12.196 3.56 0.1004 123.770 −8.16   Theo.34
Hf2SnC 3.330 14.370 4.32 0.0879 138.006   Theo.35
Hf2TlC 3.370 14.847 4.41 0.0825 146.025   Theo.33


3.2 Compound's stability

Since Hf2AC (A = Cl, Br) are newly identified compounds with unexplored properties that have not yet been experimentally synthesized, our objective is to establish their theoretical foundation through density functional theory (DFT) calculations. This theoretical investigation aims to guide future experimental and computational studies. While we have predicted the stability of Hf2ClC and Hf2BrC using first-principles calculations, additional theoretical analyses may be necessary. Therefore, we conduct a comprehensive stability assessment of the Hf2AC (A = Cl, Br) phases to provide insights for future experimental validation.

To evaluate the structural stability of Hf2AC (A = Cl, Br), we examined the c/a ratio. For M2AX-type phases, the ideal c/a value is approximately 4.89 (or 2√6).7,38 As shown in Table 1, the computed c/a values for Hf2ClC (3.77) and Hf2BrC (3.84) are reasonably close to this ideal value, indicating favorable structural stability.

Furthermore, in MAX phases, the spacing between the M and X (X = C, N) layers is characterized by the internal structural parameter zm, which ideally takes a value of approximately 0.083 (or 1/12).7,38 Our calculations reveal that the zm values for Hf2ClC and Hf2BrC are close to this ideal value, indicating minimal deviation and further supporting structural stability (see Table 1).

To further assess stability, we analyzed distortions in the unit cell using the octahedral site (Or) and trigonal prism site (Pr) parameters, defined as:5

 
image file: d5ra07272g-t1.tif(1)

Ideally, both Or and Pr should be equal to 1, with values greater or less than 1, indicating structural distortions. The Or and Pr values for Hf2ClC and Hf2BrC are quite similar, with only slight differences. The slightly higher Or and Pr values for Hf2ClC compared to Hf2BrC can be attributed to variations in atomic size, bonding characteristics, and structural distortions introduced by the halogen element (Cl vs. Br). Since Cl has a smaller atomic radius than Br, the interatomic spacing in the Hf2ClC structure is slightly reduced compared to Hf2BrC, leading to greater structural distortion. In contrast, the larger Br atoms in Hf2BrC result in a more relaxed lattice structure with slightly reduced distortions, leading to lower Or and Pr values. However, since these differences are small, both compounds remain structurally stable.

To evaluate their thermodynamic stability, we calculated the formation enthalpy (ΔHf) and binding energy (ΔEb), both of which serve as key indicators of stability. The formation enthalpy was computed using the following equations:39

 
image file: d5ra07272g-t2.tif(2)
where, Etot(Phase) represents the total energy of the Hf2AC (A = Cl, Br) phase, while EHf, EA, and EC denote the atomic energies of Hf, A (Cl or Br), and C, respectively. The variables l, m, and n represent the number of atoms in the compound's empirical formula, while N indicates the total number of atoms in the formula unit. The calculated values of ΔHf presented in Table 1, confirm the thermodynamic stability of both Hf2ClC and Hf2BrC. Moreover, similar negative energy values have been reported for other MAX phases.4,5,7,40–42 This consistency suggests that Hf2AC (A = Cl, Br) phase in our study is not only energetically favorable and exothermic but also highly stable.

Although the formation enthalpy (ΔHf) provide insights into thermodynamic stability, it does not fully capture the stability of a compound. It is also essential to examine phase stability with respect to competing phases. This approach is particularly important in materials science, chemistry, and engineering, where real-world applications depend on the compound's resistance to decomposition into other stable phases. To address this, we analyzed the phase stability by comparing the total energies of Hf2ClC and Hf2BrC with those of relevant competing phases, as obtained from the Open Quantum Materials Database (OQMD).43 For both compounds, the same set of competing phases was considered, although their compositional contributions differ slightly as dictated by the overall stoichiometry. The decomposition reactions into competing stable phases are expressed as follows:

• Hf2ClC → 0.5 × Hf2CCl2 + 0.1 × Hf6C5 + 0.4 × Hf

• Hf2BrC → 0.5 × Hf2CBr2 + 0.1 × Hf6C5 + 0.4 × Hf

These competing phases were selected based on thermodynamic stability criteria and known structural feasibility. The calculated formation energies with respect to these phases are listed in Table 1. The negative values of the formation energies (ΔEf in eV) confirm the relative stability of Hf2ClC and Hf2BrC, reinforcing the conclusions drawn from the formation enthalpy and binding energy analyses.

To further confirm the dynamical stability of Hf2AC (A = Cl, Br) MAX phases, we conducted a phonon dispersion analysis at zero pressure and temperature. This evaluation is essential for determining the feasibility of synthesizing these compounds in a hexagonal structure. Using density functional perturbation theory (DFPT), we computed the phonon spectrum dispersion.31

The phonon dispersion curves (PDCs) with total density of states, shown in Fig. 2(a and b), offer vital information about stability. A fundamental criterion for dynamical stability is the absence of imaginary phonon frequencies in the dispersion curves. Since our results show no such imaginary frequencies, we confirm that Hf2ClC and Hf2BrC are dynamically stable and can potentially be synthesized experimentally.


image file: d5ra07272g-f2.tif
Fig. 2 Calculated phonon dispersion spectra and the right panel shows the phonon density of states for (a) Hf2ClC, and (b) Hf2BrC.

Each graph displays three acoustic branches (red curves in the low-frequency region) and twenty-one optical branches (blue, green, and orange curves in the higher-frequency region). The longitudinal acoustic (LA) mode is seen progressing smoothly, while the transverse acoustic (TA) modes exhibit slight dispersion near the Γ-point. The transverse optical (TO) and longitudinal optical (LO) modes are located at higher frequencies for Hf2ClC (∼20 THz, ∼15.35 THz) compared to Hf2BrC (∼19.22 THz, ∼15.0 THz). This frequency reduction in Hf2BrC is due to Br being heavier than Cl, which lowers the vibrational frequencies as per the inverse mass relation in lattice dynamics (i.e., image file: d5ra07272g-t3.tif). As a result, Hf2BrC has softer optical phonon modes, leading to slightly lower phonon energy compared to Hf2ClC. The low-frequency acoustic phonons (red curves), responsible for lattice vibrations and heat conduction, are slightly more dispersed in Hf2BrC compared to Hf2ClC. Since Br is larger than Cl, it weakens the bonding strength in Hf2BrC, making the structure slightly more flexible. This results in lower acoustic phonon velocities, which could impact thermal conductivity. In both cases, the optical and acoustic phonon branches remain well-separated, which indicates limited phonon scattering. However, Br induces softer phonon modes compared to Cl. This may enhance acoustic–optical phonon interactions, leading to increased phonon scattering and reduced thermal conductivity. This suggests that Hf2BrC might exhibit lower thermal conductivity than Hf2ClC.

Hf2AC (A = Cl, Br) MAX phases satisfy all essential stability criteria, including formation energy, phonon dispersion curves, and mechanical stability conditions (Section 3.3). These results confirm their thermodynamic and dynamical stability, making them strong candidates for experimental synthesis. Several MAX phases, including carbides, nitrides, and borides, have already been successfully synthesized.41 Additionally, various Hf-based MAX phases, such as Hf2AlC, Hf3AlC2, Hf5Al2C3, Hf2SnC, and Hf3PbC2 have been experimentally realized.36,44–46 Given this trend, there is a strong potential for the successful synthesis of Hf2ClC and Hf2BrC, further expanding the family of Hf-based MAX phases.

3.3 Mechanical and elastic properties

Elastic constants are essential for evaluating the mechanical stability and stiffness of materials under external stress. The mechanical behavior of a crystal lattice is characterized by its second-order elastic stiffness constants, which play a crucial role in determining the mechanical stability of a material.47 These constants are derived from the strain-energy relationship as:
 
image file: d5ra07272g-t4.tif(3)
where E represents the energy of the crystal, V0 is its equilibrium volume, and ε is the applied strain.

In hexagonal crystal structures, the stiffness matrix consists of 21 components, of which only six are independent: C11, C12, C13, C33, C44, and C66. Among these, C66 is dependent and follows the relation C66 = (C11C12)/2. The computed values of these elastic constants for Hf2AC (A = Cl, Br) are presented in Table 2. To confirm mechanical stability, these values must satisfy the Born–Huang stability criteria for hexagonal structures:47

 
C11 > |C12|, 2C132 < C33(C11 + C12), C44 > 0 and C66 > 0 (4)

Table 2 The calculated stiffness constants (Cij in GPa), Cauchy pressure (C′′ in GPa), and Kleinman parameter (ζ) for Hf2AC (A = Cl, Br) MAX phases along with some other theoretical results
Compounds C11 C12 C13 C33 C44 C66 C′′ ζ Ref.
Hf2ClC 165 56 75 140 27 54 29 0.48 This work
Hf2BrC 156 78 68 147 22 39 56 0.63
Hf2AlC 329 70 81 272 126 129 −56 Ref. 23
Hf2SC 320 95 121 330 149 113 −54 Ref. 34
Hf2SnC 293 87 101 278 110 103 −23 Ref. 35
Hf2TlC 278 69 62 230 71 104 −2 Ref. 33


The computed values for Hf2ClC and Hf2BrC meet these criteria, indicating that they are mechanically stable. These results suggest that these MAX phases can be considered for further experimental exploration. Since no prior experimental or theoretical data exist for the elastic constants of Hf2ClC and Hf2BrC, a comparative analysis with other Hf-based MAX phases provides valuable insights into their mechanical behavior.

A comparison of the stiffness constants in Table 2 reveals that Hf2ClC and Hf2BrC exhibit lower mechanical stiffness than other Hf-based MAX phases. The values of C11, which represent resistance to uniaxial deformation, are 165 GPa and 156 GPa for Hf2ClC and Hf2BrC, respectively. These values are significantly lower than those of other Hf-based MAX phases, indicating that halogen-based MAX phases are more flexible. Similarly, the values of C33, which represent resistance to compression along the c-axis, are 140 GPa for Hf2ClC and 147 GPa for Hf2BrC, compared to much higher values in Hf2AlC, Hf2SC, Hf2SnC, and Hf2TlC. This reduction in stiffness along the c-axis suggests that halogen substitution weakens bonding in this direction, making these structures more compliant under mechanical stress. The shear modulus, represented by C44 and C66, is also significantly lower in Hf2ClC (27 GPa, 54 GPa) and Hf2BrC (22 GPa, 39 GPa) compared to other Hf-based MAX phases as mentioned in Table 2. This suggests that Hf2ClC and Hf2BrC have reduced resistance to shear deformation, with Hf2BrC being even more flexible due to the larger atomic size and lower electronegativity of Br, which weakens the overall bonding strength.

The stiffness constants Cij reveal the elastic anisotropy of Hf2ClC and Hf2BrC. The difference between C11 and C33 indicates stronger bonding along the a(b)-axis than the c-axis, making compression along the a-axis more difficult. The higher C11 value confirms that bonding along [100] is stronger than along [001], similar to other Hf-based MAX phases.23,33–35 The lower C44 values suggest shear deformation is easier than linear compression, providing flexibility while maintaining structural integrity. The lower values of C12 and C13 compared to C11 and C33 indicate that stress along the a-axis significantly affects mechanical properties. The equation C11 + C12 > C33 shows that the bonds are stronger bonding in the (001) plane than along the c-axis. The difference between two specific elastic constants, known as the Cauchy pressure, is expressed as C′′ = (C12C44) and serves as an indicator of a material's ductility or brittleness. A positive Cauchy pressure suggests ductility, while a negative value indicates brittleness. It also helps us understand how atoms bond in solids. A positive Cauchy pressure implies predominant ionic bonding, whereas a negative value suggests covalent bonding.48 The positive C′′ values for Hf2ClC (29 GPa) and Hf2BrC (56 GPa) (see in Table 2) confirm their ductile nature, which enhances their resistance to mechanical failure under stress. The presence of ionic bonding improves their flexibility, making them suitable for applications requiring high toughness, such as wear-resistant and impact-resistant coatings. Their ductile behavior also ensures better machinability, enabling their use in cutting tools and structural components exposed to cyclic loading. Additionally, their combination of ductility and moderate stiffness makes them promising for thermal barrier coatings, where resistance to thermal shock and mechanical integrity are crucial.

The substitution of Cl and Br in A site of Hf2AC (A = Cl, Br) significantly affects the Kleinman parameter image file: d5ra07272g-t5.tif, which describes the relative contributions of bond bending and bond stretching in response to mechanical deformation.49 The Kleinman parameter is a dimensionless parameter that typically lies in the range 0 ≤ ζ ≤ 1. The calculated ζ values increase from 0.48 in Hf2ClC to 0.63 in Hf2BrC, indicating a shift toward greater bond bending dominance. This change is attributed to the larger atomic radius of Br, which weakens Hf–Br bonds compared to Hf–Cl, making it easier for bonds to bend rather than stretch under stress. A higher ζ value in Hf2BrC suggests greater flexibility and the ability to absorb mechanical strain more effectively. This enhances its resistance to damage, making it suitable for applications requiring impact resistance and adaptability to thermal expansion. In contrast, the lower ζ in Hf2ClC indicates a more balanced contribution of bond bending and stretching, resulting in higher stiffness and structural stability. This makes Hf2ClC a better candidate for applications where mechanical strength and rigidity are prioritized over flexibility.

The polycrystalline elastic moduli, including the bulk modulus (B), shear modulus (G), Young's modulus (E), and Poisson's ratio (ν), were determined from the single-crystal elastic constants using the Voigt–Reuss–Hill (VRH) approximations.50 The Voigt and Reuss equations provide the upper and lower bounds for these moduli, and Hill's approximation takes their average to define the polycrystalline properties. The bulk modulus is calculated as BH = (BV + BR)/2, where BV and BR are the Voigt and Reuss bulk moduli, respectively. Similarly, the shear modulus is given by GH = (GV + GR)/2.

Using these values, Young's modulus and Poisson's ratio are derived from the relationships E = 9BHGH/(3BH + GH) and v = 3BH − 2 GH/2(BH + GH). The computed values for Hf2ClC and Hf2BrC, along with other MAX phases, are presented in Table 3 and shown in Fig. 3(b).

Table 3 The calculated values of polycrystalline elastic moduli (B, G, Y in GPa), Pugh's ratio (B/G), Poisson's ratio (ν), hardness parameters (Hmicro and Hmacro in GPa) and machinability index (μM) of Hf2AC (A = Cl, Br) phases along with other theoretical data
Phases B G E B/G ν Hmicro Hmacro μM Ref.
Hf2ClC 97.90 37.51 99.78 2.61 0.33 4.24 2.43 3.62 This work
110.86 45.24 119.48 2.45 0.32 5.42 Ref. 2
Hf2BrC 98.62 31.44 85.27 3.13 0.35 3.02 0.95 4.52 This work
116.38 29.45 81.47 3.95 0.38 2.29 Ref. 2
Hf2AlC 154 122 297 0.79 0.19 Ref. 23
Hf2SC 183 123 300 1.49 0.23 Ref. 34
Hf2SnC 161 102 252 1.58 0.24 Ref. 35
Hf2TlC 129 87 214 1.48 0.22 Ref. 33



image file: d5ra07272g-f3.tif
Fig. 3 Comparison of (a) elastic constants and (b) elastic moduli of Hf2ClC and Hf2BrC with other Hf-based MAX compounds.

The bulk modulus (B) represents a material's resistance to uniform compression. From Fig. 3(b), it can be observed that Hf2ClC (97.90 GPa) and Hf2BrC (98.62 GPa) exhibit significantly lower bulk modulus (B) values compared to Hf2AlC, Hf2SC, Hf2SnC and Hf2TlC. This indicates that halogen-based MAX phases are more compressible and flexible. The lower B values suggest weaker interatomic bonding in Hf2ClC and Hf2BrC, making them more deformable under pressure. The bulk modulus (B) shows a slight increase of 0.73% upon Cl-to-Br substitution, indicating a minimal change in compressibility. This suggests that replacing Cl with the larger Br atom does not significantly alter the overall resistance to volumetric deformation. Despite this minor variation, both Hf2ClC and Hf2BrC remain highly compressible compared to other Hf-based MAX phases, reinforcing their potential for applications where flexibility and adaptability under mechanical stress are required.

The shear modulus (G), which measures resistance to shape deformation, is also lower in Hf2ClC (37.51 GPa) and Hf2BrC (31.44 GPa) compared to other Hf-based MAX phase.23,33–35 This suggests that halogen-containing MAX phases have reduced resistance to shear forces, making them more ductile and less rigid than other Hf-based MAX phases. The value of G decreases significantly by 16.16% upon Cl-to-Br substitution, indicating that Hf2BrC becomes more flexible and less resistant to shear deformation. This reduction in shear modulus can be attributed to the larger atomic radius and lower electronegativity of Br compared to Cl, which weakens interatomic bonding and reduces resistance to shear forces. The lower shear modulus enhances the material's ductility, making Hf2BrC more adaptable under mechanical stress. The improved ductility makes Hf2BrC suitable for environments subjected to mechanical shocks or vibrations, such as aerospace components and thermal barrier coatings.

Young's modulus (E), which defines the stiffness of a material, follows the same trend. This confirms that Hf2ClC and Hf2BrC are mechanically softer and more flexible compared to these Hf-based MAX phases.

The value of E decreases by 14.57% upon Cl-to-Br substitution, confirming reduced stiffness and structural rigidity in Hf2BrC. This decline is primarily due to the weaker bonding strength introduced by the larger Br atoms, which reduce the overall bond strength and increase lattice flexibility. This reduction in stiffness enhances the material's toughness and damage tolerance, making Hf2BrC a better candidate for applications requiring impact resistance and flexibility, such as protective coatings and shock-absorbing layers. The decreased E also suggests improved machinability, making it easier to process and shape for engineering applications.

In addition, Young's modulus (E) plays a crucial role in determining thermal shock resistance (R), as seen in the equation image file: d5ra07272g-t6.tif, where σb stands for strength and α for thermal expansion coefficient.51 A higher Young's modulus leads to a lower thermal shock resistance, meaning materials with greater stiffness are more susceptible to thermal stress fractures. Since Hf2BrC has a lower Young's modulus (85.27 GPa) compared to Hf2ClC (99.78 GPa), it is expected to have better thermal shock resistance. This is because lower stiffness allows the material to absorb thermal stress more effectively, reducing the likelihood of crack propagation under rapid temperature changes. The greater thermal shock resistance of Hf2BrC makes it a better candidate for these applications, as it can withstand sudden temperature variations without failure. On the other hand, Hf2ClC, with its higher Young's modulus, may be more suitable for applications where structural rigidity and mechanical strength are prioritized over thermal shock resistance, such as load-bearing components in high-performance machinery.

The B/G ratio (Pugh's ratio) and Poisson's ratio (ν) are key indicators of a material's mechanical properties, particularly its brittleness or ductility.52,53 The B/G ratio and Poisson's ratio (ν) of Hf2ClC and Hf2BrC are greater than those of other Hf-based MAX phases, as mentioned in Table 3. Generally, if B/G > 1.75, the material tends to be ductile, and if B/G < 1.75, it is more likely to be brittle. Both Hf2ClC and Hf2BrC have B/G ratios above 1.75, indicating they are ductile materials. However, Hf2BrC shows a higher B/G ratio (3.13) compared to Hf2ClC (2.61), suggesting it is slightly more ductile. This change is attributed to the larger ionic radius of Br compared to Cl, which allows for more flexible bonding and less rigidity in the structure, enhancing ductility.

Poisson's ratio (ν) reveals the compressibility, bonding forces, and nature of the material. The value of v greater than 0.26 typically suggests ductility, while a value below 0.26 tends to indicate brittleness. Both values are above 0.26, indicating that both materials are ductile. The substitution of Cl with Br increases the Poisson's ratio slightly, which aligns with the trend seen in the B/G ratio and Cauchy pressure. This implies that the presence of Br in Hf2BrC results in a slightly higher ductility compared to Hf2ClC. This characteristic could make Hf2BrC suitable for use in industries requiring durable, high-performance.

Moreover, the Poisson's ratio (ν) of Hf2AC (A = Cl, Br) is critical in understanding the nature of the interatomic forces and the chemical bonding in these compounds. The values of ν for Hf2ClC and Hf2BrC are 0.33 and 0.35, respectively, both of which fall within the range of 0.25 to 0.50, indicating that the interatomic forces in these materials are central in nature.49 This suggests that the material's deformation behavior is influenced by central forces between the atoms, rather than directional covalent forces.

The Poisson's ratio also gives an explanation for the bonding character. For purely covalent materials,53 Poisson's ratio is typically around 0.10, while for purely ionic materials, it tends to be closer to 0.25. The calculated Poisson's ratios for Hf2ClC and Hf2BrC (both above 0.25) suggest that these compounds exhibit a significant ionic contribution to their chemical bonding. When considering the substitution of Cl with Br, the increase in Poisson's ratio from 0.33 to 0.35 (Hf2ClC to Hf2BrC) indicates a slight enhancement in the ionic bonding character. Br, being larger and less electronegative than Cl, may result in a more flexible atomic structure, which could lead to a slightly greater ionic interaction. This subtle increase in ionic bonding may contribute to Hf2BrC's slightly greater ductility compared to Hf2ClC, as reflected in its higher B/G ratio. Thus, the Cl-to-Br substitution slightly enhances the ionic character of the bonding in Hf2BrC, which could improve its mechanical properties, making it more ductile.

The hardness of a material is a key factor in its ability to resist deformation, wear, and indentation under force. For Hf2AC (A = Cl, Br) compounds, the hardness parameters (Hmicro, Hmacro) are calculated by following relation:54,55

 
image file: d5ra07272g-t7.tif(5)
 
Hmacro = 2(k2G)0.585 − 3 (6)
where, k and G refers to Pugh's ratio (k = G/B) and shear modulus, respectively.

The calculated hardness values for Hf2ClC (Hmicro = 4.24 GPa, Hmacro = 2.43 GPa) and Hf2BrC (Hmacro = 3.02 GPa, Hmicro = 0.95 GPa) show that both compounds are relatively soft, with Hf2BrC being softer than Hf2ClC (see Table 3). The Cl-to-Br substitution in Hf2AC (A = Cl, Br) leads to a reduction in hardness, with Hf2BrC showing significantly lower hardness compared to Hf2ClC. The lower hardness of Hf2BrC can be attributed to the larger atomic size of Br compared to Cl, which likely results in weaker atomic bonds and a more easily deformable structure. The relatively soft nature of Hf2ClC and Hf2BrC makes these materials suitable for applications where high hardness is not a critical requirement, but good ductility and flexibility are desired. Additionally, Hf2BrC, with its lower hardness, may find use in applications that require easier machining or shaping, such as in electronic components or catalysts where moderate mechanical strength suffices.

The machinability index (μM = B/C44) is an important parameter for assessing the plasticity and lubricating properties of a material.56 It is influenced by factors such as elastic properties, hardness, and the ability of a material to undergo plastic deformation. A higher μM value indicates improved lubrication, reduced friction, and enhanced machinability, making the material easier to shape and process. The calculated values of μM for Hf2ClC and Hf2BrC are presented in Table 3, showing that Hf2BrC has better machinability compared to Hf2ClC. This is due to Br's larger atomic size, which weakens atomic bonding, increasing plasticity and easing deformation. Hardness and machinability are inversely related, meaning materials with lower hardness are generally easier to machine. The decrease in hardness for Hf2BrC is 28.77% (microhardness) and 60.91% (macro-hardness), while its machinability index is 24.86% higher than Hf2ClC. This confirms that as hardness decreases, machinability improves due to weaker atomic bonding, allowing for easier deformation. The higher machinability of Hf2BrC makes it suitable for high-speed machining, reducing tool wear and processing costs.

3.4 Elastic anisotropy

In this study, we explore the elastic anisotropy of Hf2AC (A = Cl, Br), which is crucial for understanding its mechanical behavior under various conditions. Elastic anisotropy influences plastic deformation, crack propagation, and stress tolerance. The crystal structure (see Fig. 1) reveals distinct atomic arrangements along the a(b) and c directions, leading to differences in bond strength. The stiffness along a(b) is represented by C11, while the stiffness along c is denoted by C33. Since C11C33, this confirms anisotropy in mechanical properties. The shear anisotropic factors (A1, A2, A3) provide insight into the bonding strength variations across different atomic planes. The shear anisotropy factor for {100} shear planes between [011] and [010] directions is,57
 
image file: d5ra07272g-t8.tif(7)
for {010} shear planes between [101] and [001] directions
 
image file: d5ra07272g-t9.tif(8)
for {001} shear planes between [110] and [010] directions
 
image file: d5ra07272g-t10.tif(9)

The calculated shear anisotropic factors of Hf2AC (A = Cl, Br) are listed in Table 4. The calculated anisotropy values indicate that both Hf2ClC and Hf2BrC exhibit moderate anisotropy, with deviations from the ideal isotropic value (1). However, Cl-to-Br substitution significantly impacts the degree of anisotropy. For Hf2BrC, A1 and A2 values decrease compared to Hf2ClC, dropping from 0.69 to 0.52, indicating an increase in anisotropy in {100} and {010} shear planes. This suggests that Br's larger atomic size weakens directional bonding, leading to increased deformation along these planes. Despite these changes, A3 remains 1.0 for both compounds, confirming isotropy in {001} plane. This comparison effectively highlights the practical implications of elastic anisotropy in Hf2ClC and Hf2BrC. Hf2BrC's increased anisotropy suggests greater plastic deformation, making it more adaptable for flexible structural applications where some deformation can be tolerated. Conversely, Hf2ClC's more uniform bonding structure enhances wear resistance and fracture toughness, making it more suitable for high-stress environments such as protective coatings or structural components.

Table 4 The calculated shear anisotropic factors (A1, A2 and A3), percentage anisotropy in the compression and shear (AB and AG), universal anisotropic index (AU), equivalent Zener anisotropic factor (Aeq), compressibility ratio (kc/ka), and universal log-Euclidean index (AL) for studied compounds Hf2AC (A = Cl, Br)
Compounds A1 A2 A3 AB AG AU Aeq kc/ka AL Ref.
Hf2ClC 0.69 0.69 1.0 0.0 0.049 0.52 1.91 2.38 1.27 This work
Hf2BrC 0.52 0.52 1.0 0.0 0.045 0.48 1.86 2.12 1.08


We have also calculated the percentage elastic anisotropy in shear (AG) and compression (AB), the universal anisotropy factor (AU), the equivalent Zener anisotropy factor (Aeq) by the following common relation58,59 and their estimated values are displayed in Table 4.

 
AB = (BVBR)/(BV + BR) (10)
 
AG = (GVGR) + (GV + GR) (11)
 
image file: d5ra07272g-t11.tif(12)
 
image file: d5ra07272g-t12.tif(13)

For an isotropic crystal, AB = AG = 0. However, the degree of anisotropy is represented by any variation larger than zero. For both materials, the percent elastic anisotropy in compression (AB) is 0.0, meaning they are isotropic in compressibility, implying uniform bulk modulus distribution. However, the percentage elastic anisotropy in shear (AG) is 0.049 for Hf2ClC and 0.045 for Hf2BrC, confirming slight anisotropy in shear deformation. The lower AG value for Hf2BrC suggests that Br substitution weakens the directional dependency of shear modulus, slightly reducing anisotropy in shear compared to Cl.

Ranganathan and Ostoja-Starzewski59 have introduced the concept of the universal elastic anisotropic index (AU) which is a widely used index for determining anisotropy due to its capacity to be applied to all potential crystal symmetries. The crystal is completely isotropic when the value of the universal anisotropy index (AU) is exactly equal to zero (AU = 0). The universal anisotropy factor (AU) values of 0.52 (Hf2ClC) and 0.48 (Hf2BrC) further indicate that both compounds deviate from perfect isotropy. Hf2BrC has a slightly lower AU value, signifying that the introduction of Br reduces overall elastic anisotropy.

For an isotropic crystal, the equivalent Zener anisotropy factor is equal to one (Aeq = 1). The equivalent Zener anisotropy factor (Aeq) values of 1.91 (Hf2ClC) and 1.86 (Hf2BrC) support this trend, confirming moderate anisotropy. Br substitution weakens atomic bonding, which makes the material less stiff in one direction and a little more flexible in another. This makes Hf2BrC more adaptable for applications requiring flexibility, while Hf2ClC remains preferable for high-stress applications requiring superior mechanical stability.

The compressibility ratio image file: d5ra07272g-t13.tif provides insight into how a material deforms under stress along different crystallographic directions.60 A higher ratio indicates greater compressibility along the c-axis compared to the a-axis, signifying directional anisotropy in response to external forces. For Hf2ClC, kc/ka is 2.38, while for Hf2BrC, it is 2.12 (see Table 4). This suggests that Hf2ClC exhibits greater compressibility along the c-axis, making it more adaptable to deformation in this direction under applied stress. The substitution of Cl with Br in Hf2AC (A = Cl, Br) leads to a decrease in the compressibility ratio (kc/ka) from 2.38 to 2.12, indicating an 11% reduction. This suggests that Hf2BrC has lower anisotropy in compressibility compared to Hf2ClC. Br is larger than Cl, so it doesn't weaken the bonding along the c-axis as much as Cl does. This makes the compressibility between the a- and c-axes more even. As a result, Hf2BrC exhibits a more balanced mechanical response and improved structural stability under stress. On the other hand, Hf2ClC, with its higher kc/ka ratio, is more compressible along the c-axis, making it more flexible and adaptable to deformation in this direction.

We have also calculated the universal log-Euclidean index (AL) can be defined using the log-Euclidean formula.58

 
image file: d5ra07272g-t14.tif(14)
where, the subscripts V and R represent the Voigt and Reuss approximations, respectively. For the universal log-Euclidean index calculation, the Voigt and Reuss values of C44 are obtained from ref. 58 which are CV44 = [CR44 + (3/5)((C11C12 − 2C44)2/(3(C11C12) + 4C44))], and CR44 = [(5/3)((C44(C11C12))/3(C11C12) + 4C44)].

The universal log-Euclidean index (AL) for Hf2ClC and Hf2BrC is 1.27 and 1.08, respectively, indicating a 15% reduction with Br substitution. Since most crystalline solids have AL < 1.0, both compounds exhibit significant anisotropy, with Hf2ClC being more anisotropic. The higher AL and compressibility ratio (kc/ka) in Hf2ClC (2.38) compared to Hf2BrC (2.12) reflects its stronger layered character and pronounced anisotropic mechanical behavior. Br substitution weakens interlayer bonding due to its larger atomic radius, leading to a more uniform elastic response and reduced anisotropy in Hf2BrC. This translates to enhanced structural flexibility and adaptability, making Hf2BrC suitable for applications requiring moderate mechanical compliance. In contrast, Hf2ClC, with its higher anisotropy and directional bonding, is suitable for high-stress environments where mechanical rigidity is crucial.

Thus, replacing Cl with Br reduces anisotropy, balancing structural properties and making Hf2BrC a more isotropic material with improved mechanical versatility.

3.5 Thermo-mechanical properties

In the study of acoustics for crystalline solids, sound velocity is a critical parameter, especially in elastically anisotropic materials where sound velocities vary depending on the direction in which elastic waves propagate. These materials exhibit different velocities for longitudinal and transverse waves, which are key in understanding their mechanical, electrical, and thermal conductivities. The sound velocity is directly related to the material's elastic moduli and can be calculated using the longitudinal (vl) and transverse (vt) sound velocities. The average sound velocity, va, is derived from the following formula:61
 
image file: d5ra07272g-t15.tif(15)
where the transverse sound velocity vt is given by:
 
image file: d5ra07272g-t16.tif(16)
And the longitudinal sound velocity vl is:
 
image file: d5ra07272g-t17.tif(17)

Acoustic impedance (Z) of a medium is another important parameter that can be calculated from its density (ρ) and shear modulus (G):49

 
image file: d5ra07272g-t18.tif(18)

Higher acoustic impedance values are observed in denser and stiffer materials. The unit of acoustic impedance is Rayl (1 Rayl = 1 kg m−2 s−1 = 1 Ns m3).

Finally, the radiation intensity (I) is also an important acoustical factor. It is proportional to the surface velocity of the material and scales with its density and modulus of rigidity. The radiation intensity can be expressed as:49

 
image file: d5ra07272g-t19.tif(19)

The calculated values of sound velocities, acoustic impedance, and radiation factor for Hf2AC (A = Cl, Br) are listed in Table 5 along with some other theoretical results. The substitution of Cl with Br in Hf2AC (A = Cl, Br) significantly affects its acoustic properties due to the difference in atomic mass and bonding strength. The mass density increases from 10.17 g cm−3 in Hf2ClC to 10.53 g cm−3 in Hf2BrC, primarily due to the larger atomic weight of Br. This increase in density leads to a noticeable reduction in sound velocities. The transverse sound velocity decreases from 1921 m s−1 in Hf2ClC to 1729 m s−1 in Hf2BrC, reflecting a 10% reduction, while the longitudinal sound velocity drops from 3814 m s−1 to 3654 m s−1, indicating a 4.2% decrease. The average sound velocity also declines by approximately 9.7%, from 2154 m s−1 to 1945 m s−1, which suggests that the propagation of sound waves through the material becomes slower with Br substitution. Materials with lower acoustic radiation and slower phonon propagation tend to have lower thermal conductivity, making them suitable for thermal insulation or thermal barrier coatings. Additionally, because they minimize energy loss from lattice vibrations, they can enhance the efficiency and durability of devices operating at high temperatures or under thermal stress. Substituting Br for Cl increases density, slows sound propagation, and reduces acoustic radiation, making Hf2BrC better at damping vibrations and insulating heat, which is advantageous for high-temperature and thermal barrier applications.

Table 5 The calculated mass density (ρ in gm cm−3), sound velocities (vt, vl and va in ms−1), acoustic impedance (Z in Rayl), radiation factor (I in m4 kg−1 s−1) for Hf2AC (A = Cl, Br)
Compounds ρ vt vl va Z (106) I Ref.
Hf2ClC 10.17 1921 3814 2154 19.53 0.19 This work
Hf2BrC 10.53 1729 3654 1945 18.21 0.16
Hf2AlC 9.65 3560 5730 3920 Ref. 23
Hf2SC 11.02 3840 5300 3740 Ref. 34
Hf2TlC 13.04 2589 4345 2866 Ref. 33


Furthermore, the acoustic impedance, which measures a material's ability to transmit sound, decreases from 19.53 × 106 Rayl in Hf2ClC to 18.21 × 106 Rayl in Hf2BrC, showing a 6.8% reduction. This suggests that Hf2BrC exhibits weaker impedance matching, making it less effective in transmitting sound. Additionally, the radiation factor, which indicates a material's ability to radiate sound energy efficiently, decreases by 15.8%, from 0.19 to 0.16 m4 kg−1 s−1. The reduction in both acoustic impedance and radiation efficiency implies that Hf2BrC has weaker bonding interactions and a lower capacity to propagate and radiate sound compared to Hf2ClC. Overall, Cl-to-Br substitution results in a material with reduced sound transmission and increased structural flexibility, making Hf2BrC more suitable for applications requiring lower acoustic impedance, while Hf2ClC remains preferable for high-speed acoustic applications.

In conclusion, the incorporation of halogens like Cl and Br in Hf2AC compounds leads to a reduction in sound velocities and mechanical stiffness compared to other hafnium-based materials, as shown in Table 5. This suggests that these halogen-substituted compounds may exhibit more flexibility or adaptability under certain conditions.

3.6 Thermo-lattice properties

The Debye temperature (ΘD) is a crucial thermophysical parameter that provides significant insights into the thermal and mechanical behavior of materials. It is related to the material's lattice vibrations, thermal conductivity, interatomic bonding, and other properties such as melting temperature and specific heat capacity. The Debye temperature effectively serves as a cutoff for the phonon energy distribution and helps distinguish between the classical and quantum mechanical behavior of phonons. The values of ΘD are obtained from the following equation:62
 
image file: d5ra07272g-t20.tif(20)
where h is Planck's constant, n is the number of atoms in the unit cell, kB is the Boltzmann constant, and V0 is the equilibrium volume. The Debye temperature (ΘD) for Hf2ClC is 250 K, while for Hf2BrC, it is 221 K are listed in Table 6. This shows a reduction of approximately 11.6% in the Debye temperature when Cl is replaced with Br. The decrease in the Debye temperature suggests that the interatomic bonding in Hf2BrC is weaker than in Hf2ClC. This results in less efficient lattice vibrations and lower mechanical wave velocities in Hf2BrC compared to Hf2ClC. The reduced Debye temperature in Hf2BrC indicates that the material has a lower thermal conductivity due to weaker atomic interactions. This behavior can be attributed to the larger atomic radius of Br compared to Cl, which introduces more spacing between the atoms, weakening the bonding and lowering the lattice thermal conductivity. These weaker atomic interactions in Hf2BrC result in a lower Debye temperature and reduced thermal conductivity. In contrast, Hf2ClC, with its stronger atomic interactions, is expected to exhibit a higher lattice thermal conductivity. The stronger bonds in Hf2ClC allow for more efficient heat transfer through the crystal lattice. Therefore, the substitution of Cl with Br reduces the atomic bonding strength, leading to lower thermal conductivity in Hf2BrC compared to Hf2ClC.
Table 6 The calculated values of Debye temperature (ΘD in K), melting temperature (Tm in K), thermal expansion coefficient (α in K−1), heat capacity (ρCP in J m−3 K−1), Grüneisen parameter (γ), dominant phonon wavelength (λdom in m), minimum thermal conductivity (kmin in W m−1 K−1) and lattice thermal conductivity (kph in W m−1 K−1 at 300 K) for Hf2AC (A = Cl, Br)
Compound ΘD Tm γ kph kmin α (×10−5) ρCP (×106) λdom (×10−12)
Hf2ClC 250 1059 1.98 2.61 0.722 4.27 2.51 89.51
Hf2BrC 221 1044 2.18 1.65 0.630 5.09 2.33 81.10


The melting temperature63 (Tm = 354 + 1.5 (2C11 + C33)), of a material is a key thermophysical property that reflects the strength of atomic interactions and the material's thermal stability. For Hf2ClC, the calculated melting temperature is 1059 K, while for Hf2BrC, it is 1044 K. This difference indicates a 1.4% decrease in the melting temperature when Cl is substituted with Br. The lower melting temperature in Hf2BrC can be attributed to the relatively weaker atomic interactions compared to Hf2ClC. As previously discussed, Br atoms are larger than Cl atoms, which increases the interatomic spacing and weakens the bonds between atoms. The correlation between melting temperature and Young's modulus (E) also supports this observation. A higher Young's modulus typically indicates stronger atomic bonding and greater material stiffness.

The Grüneisen parameter image file: d5ra07272g-t21.tif is an essential measure of anharmonic effects in crystalline solids, influencing thermal expansion, phonon interactions, and thermomechanical properties.64 Grüneisen parameters for different polycrystalline materials are within the expected range [0.85–3.53] with the Poisson's ratio in the range of 0.05–0.46 which are good agreement with our calculated values.64 The calculated values for Hf2ClC and Hf2BrC are 1.98 and 2.18, respectively, indicating that Hf2BrC exhibits a higher degree of anharmonicity than Hf2ClC. The 10.1% increase in the Grüneisen parameter when Cl is replaced with Br suggests that the lattice vibrations in Hf2BrC are more sensitive to temperature changes and exhibit stronger anharmonic interactions. Additionally, the reduction in Debye temperature (by 11.6%) and melting temperature (by 1.4%) in Hf2BrC further supports the trend of increased anharmonicity, as lower bond strength and structural rigidity contribute to greater phonon–phonon interactions.

Lattice thermal conductivity (kph) is one of the most fundamental and crucial properties of a material, especially for high-temperature applications, and describes how efficiently heat is conducted through phonon propagation inside a crystal. For Hf2ClC and Hf2BrC compounds, lattice thermal conductivity (κph) is estimated at 300 K using Slack's formula:65

 
image file: d5ra07272g-t22.tif(21)
Here, Mav is the average atomic mass (in kg mol−1), δ is the cubic root of the average atomic volume (in units of m3), n is the number of atoms per unit cell, T is the absolute temperature, γ is the Grüneisen parameter, and A is a factor (in W-mol kg−1 m−2 K−3) dependent on γ. The coefficient A(γ) in Slack's model is determined by Julian's equation:66
 
image file: d5ra07272g-t23.tif(22)

The calculated values of lattice thermal conductivities at room temperature under study are listed in Table 6. The kph value of Hf2ClC and Hf2BrC at 300 K is 2.61 W mK−1 and 1.65 W mK−1, respectively, indicating a significant reduction of approximately 36.8% when Cl is substituted with Br. This reduction in thermal conductivity aligns with the observed decrease in Debye temperature, as lower Debye temperatures correlate with weaker interatomic bonding and reduced phonon propagation efficiency. The larger Br atomic radius increases spacing, enhancing phonon scattering and reducing lattice thermal conductivity. As a result, Hf2BrC exhibits greater thermal insulation properties compared to Hf2ClC, making it a more suitable candidate for applications requiring thermal barrier coatings and thermoelectric materials. Conversely, Hf2ClC, with its higher thermal conductivity, is better suited for heat dissipation applications in microelectronic and nanoelectronic devices.

The lattice thermal conductivity of a compound approaches a minimal value at high temperatures above the Debye temperature, which is known as minimum thermal conductivity (kmin), which is independent of further increasing temperature. The minimum thermal conductivity estimated by using the following equation:67

 
image file: d5ra07272g-t24.tif(23)
where M is molar mass, and NA is Avogadro's number.

The minimum thermal conductivity (kmin) of Hf2ClC is 0.722 W mK−1, while for Hf2BrC, it is 0.630 W mK−1, indicating a decrease of approximately 12.7% when Cl is replaced by Br. Since kmin is closely related to sound velocity and Debye temperature, the lower values in Hf2BrC suggest that it has weaker atomic interactions and a less rigid lattice compared to Hf2ClC. The minimum thermal conductivity of the studied MAX phase compounds, Hf2ClC and Hf2BrC, is significantly lower than the reference value of 1.25 W mK−1, making them suitable candidates for thermal barrier coatings (TBC) applications. The lower kmin values indicate that these materials exhibit limited heat conduction at elevated temperatures, an essential property for TBCs used in aerospace and energy applications.

The substitution of Cl with Br in Hf2AC (A = Cl, Br) significantly affects the thermal expansion coefficient, volumetric heat capacity, and dominant phonon wavelength, leading to variations in thermal performance. The calculated values of these parameters are listed in Table 6. The thermal expansion coefficient60 image file: d5ra07272g-t25.tif of Hf2BrC is higher than that of Hf2ClC, indicating that Hf2BrC undergoes greater expansion with increasing temperature. This increase is primarily attributed to the larger atomic radius of Br. Since the thermal expansion coefficient is inversely related to the melting temperature, the lower melting point of Hf2BrC further supports this observation. In practical applications, a lower thermal expansion coefficient is desirable for thermal barrier coatings to prevent thermal stress and structural instability. Therefore, Hf2ClC, with its lower expansion, is a more suitable candidate for coatings in high-temperature environments.

The volumetric heat capacity image file: d5ra07272g-t26.tif of Hf2ClC is higher than that of Hf2BrC, suggesting that it has a greater ability to store thermal energy and maintain thermal stability.49 A higher heat capacity results in lower thermal diffusivity and contributes to better thermal conductivity. Since heat capacity plays a crucial role in determining how well a material responds to temperature fluctuations, the higher value in Hf2ClC aligns with its greater lattice thermal conductivity. This characteristic makes Hf2ClC more effective in heat dissipation applications, such as microelectronics and aerospace coatings, where efficient thermal management is essential.

The dominant phonon wavelength image file: d5ra07272g-t27.tif of Hf2ClC is longer than that of Hf2BrC, which indicates a higher average sound velocity and lower phonon scattering in the former.49 A longer phonon wavelength facilitates heat transfer by reducing phonon–phonon interactions, leading to improved thermal conductivity. Conversely, the shorter phonon wavelength in Hf2BrC implies increased phonon scattering, which hinders heat conduction and reduces thermal conductivity. This characteristic makes Hf2BrC a potential candidate for thermal insulation and thermoelectric applications, where lower thermal conductivity is advantageous for energy conversion efficiency and thermal management.

Overall, the presence of Cl enhances thermal conductivity, stability, and lower expansion, making Hf2ClC preferable for applications requiring efficient heat dissipation and structural reliability. On the other hand, Br substitution increases phonon scattering and thermal expansion while reducing heat capacity, making Hf2BrC more suitable for thermal insulation and thermoelectric applications.

3.7 Electronic properties

This section provides a detailed comparison of the electronic band structures of Hf2ClC and Hf2BrC, highlighting the effect of substituting Cl with Br. The calculated band structures of Hf2ClC and Hf2BrC are shown in Fig. 4(a and b). The band structure calculations, performed using the GGA-PBE approximation, span an energy range from −4.0 to 4.0 eV, with the Fermi level set at 0 eV. The computed band structures along high-symmetry directions in the Brillouin zone indicate that both materials exhibit metallic behavior, as evidenced by multiple bands crossing the Fermi level. The dispersive nature of the band structure along different crystallographic directions provides crucial insights into the electronic transport properties of Hf2ClC and Hf2BrC. The dispersion of energy bands along the c-direction (Γ–A, H–K, and M–L) and the ab-plane (A–H, K–Γ, Γ–M, and L–H) differs significantly due to variations in bonding strength and orbital hybridization. Along the c-direction (Γ–A, H–K, and M–L), the band dispersion is relatively weaker in both compounds, indicating higher electronic localization and lower mobility along this axis. The Mulliken population analysis reveals distinct bonding characteristics in Hf2AC (A = Cl, Br) MAX phases (Table 7). Hf atoms exhibit small positive charges (+0.56e for Hf2ClC and +0.21e for Hf2BrC), while C atoms carry strongly negative charges (−0.85e), confirming substantial charge transfer from Hf to C and the presence of strong polar covalent Hf–C bonds within the layers. In contrast, the halogen atoms display different behaviors: Cl is slightly negative (−0.26e), indicating an ionic Hf–Cl interaction, whereas Br is slightly positive (+0.44e), reflecting a reversal in charge transfer. This reversal arises from Br's lower electronegativity and larger size, which weaken the ionic interaction and introduce a more covalent, polarizable Hf–Br bond. Consequently, interlayer bonding in Hf2BrC is weaker, leading to enhanced ductility, reduced stiffness, and stronger anisotropic electronic behavior compared to Hf2ClC. However, the effect of Cl and Br substitution is evident: in Hf2ClC, the bands in these directions are slightly more dispersive than in Hf2BrC, suggesting stronger interlayer coupling and better charge transport along the c-axis. Br has a larger atomic radius, which increases the space between layers. This reduces orbital overlap and makes the bands flatter, which lowers the electrical conductivity along the c-direction.
image file: d5ra07272g-f4.tif
Fig. 4 Calculated band structure of (a) Hf2ClC and (b) Hf2BrC along the high symmetry directions in the Brillouin zone at ambient conditions.
Table 7 The calculated orbital charges (electron), atomic Mulliken charges (electron), effective valence (electron), and Hirshfeld charge (electron) of the Hf2AC (A = Cl, Br) MAX compounds
Compounds Species Mulliken atomic populations Mulliken charge Formal ionic charge Effective valence Hirshfeld charge Effective valence
s p d Total
Hf2ClC C 1.53 3.33 0.0 4.85 −0.85 −4 3.15 −0.36 3.64
Cl 1.93 5.34 0.0 7.26 −0.26 −1 0.74 0.03 0.97
Hf 0.36 0.17 2.91 3.44 0.56 +4 3.44 0.17 3.83
Hf2BrC C 1.52 3.33 0.0 4.85 −0.85 −4 3.15 −0.36 3.64
Br 1.31 5.25 0.0 6.56 0.44 −1 0.56 0.06 0.94
Hf 0.43 0.47 2.90 3.79 0.21 +4 3.79 0.15 3.85


On the other hand, the bands along the ab-plane (A–H, K–Γ, Γ–M, and L–H) show stronger dispersion in both compounds. This means that charge transport in the plane is more efficient. This is due to the presence of stronger covalent interactions within the layers, particularly between Hf and C atoms. However, a noticeable difference exists between Hf2ClC and Hf2BrC: the bands in Hf2ClC are more dispersive, suggesting enhanced electronic delocalization and lower effective mass of charge carriers in the ab-plane. This higher dispersion contributes to better electronic conductivity, making Hf2ClC more suitable for electronic applications requiring high carrier mobility. In Hf2BrC, the bands in the ab-plane are slightly flatter, implying increased carrier scattering and reduced conductivity due to weaker orbital interactions. Overall, the comparison of band dispersion along the c-direction and ab-plane reveals that Hf2ClC has greater electronic transport properties, particularly within the layers, due to stronger hybridization and better orbital overlap. On the other hand, Hf2BrC exhibits more localized electronic states, leading to higher effective mass and reduced carrier mobility, especially along the c-direction. These findings highlight the anisotropic electronic nature of these materials and their potential applications, with Hf2ClC being more suitable for conductive applications and Hf2BrC potentially offering advantages in thermal barrier coatings or electronic devices where controlled conductivity is required. The anisotropic transport observed in Hf2AC (A = Cl, Br) aligns well with trends seen in other MAX phases.5,7,68,69

The density of states (DOS) plots in Fig. 5(a and b) provide deeper insight into the electronic structure of Hf2ClC and Hf2BrC, complementing the band structure analysis from Fig. 4. The variations in band dispersion along the c-direction (Γ–A, H–K, and M–L) and ab-plane (A–H, K–Γ, Γ–M, and L–H) can be directly linked to the differences in the projected DOS (PDOS) contributions of Hf, Cl, Br, and C atoms.


image file: d5ra07272g-f5.tif
Fig. 5 Representation of the total and partial density of states for (a) Hf2ClC and (b) Hf2BrC.

From Fig. 5(a and b), it is evident that the Hf-5d orbitals play a dominant role in the electronic states near the Fermi level (EF). These orbitals are mainly responsible for the dispersive nature of the bands and charge transport characteristics. In Hf2ClC, the Hf-5d states are more delocalized, contributing to the stronger dispersion observed in the band structure, particularly along the ab-plane, where covalent interactions between Hf and C are more prominent. This increased dispersion corresponds to lower effective mass and higher electronic mobility.

In contrast, in Hf2BrC, the DOS shows that the Br-4p states contribute differently compared to Cl-3p states in Hf2ClC. The Br-4p orbitals are more localized, leading to reduced hybridization with Hf-5d states. This weaker interaction results in flatter bands near the Fermi level, especially along the c-direction, where interlayer interactions are already weak. As a result, electronic conductivity is lower in Hf2BrC compared to Hf2ClC. The localization of Br states also has an effect on the dispersion along the ab-plane, but not as much as it does along the c-direction. Furthermore, the contribution of C-2p orbitals remains relatively similar in both compounds, indicating that variations in electronic dispersion are primarily influenced by the substitution of Cl with Br. The key factor in the observed differences in band dispersion is the reduced hybridization between Hf-5d and Br-4p states. Therefore, the DOS analysis confirms that the Hf-5d and halogen (Cl-3p or Br-4p) orbitals govern the band structure variations.

The electronic properties of Hf2AC (A = Cl, Br) MAX phases are strongly influenced by their total density of states (TDOS), which plays a crucial role in assessing electrical stability. The calculated N(EF) values for Hf2ClC and Hf2BrC are 4.61 and 5.19 states per eV per unit cell, respectively. Lower N(EF) values indicate greater electrical stability, suggesting that Hf2ClC is slightly more stable than Hf2BrC. These values align with the stability range (2–12 states per eV per unit cell) found in other MAX phases, confirming the robust electrical stability of both materials.70 A key feature in both compounds is the presence of a pseudo-gap near EF, which separates bonding and anti-bonding states. The pseudo-gap functions as a barrier, restricting electron transitions and enhancing stability. The Fermi level lies within the bonding region, further supporting the stability of these structures. The substitution of Cl with Br slightly reduces the pseudo-gap, which correlates with a higher N(EF), leading to a minor decrease in stability for Hf2BrC compared to Hf2ClC. This modification occurs due to Br's larger atomic radius and lower electronegativity, which weaken the hybridization between Hf and A-site atoms, making electronic states more localized and reducing charge transport efficiency.

3.8 Charge density distribution map

The electronic charge density distribution maps for Hf2ClC and Hf2BrC reveal crucial insights into their chemical bonding and charge transfer characteristics. The color gradient in Fig. 6 illustrates the variation in electronic charge, with blue representing high charge density and red indicating low charge density. The substitution of Cl with Br influences the bonding nature within these MAX phases, affecting charge accumulation and hybridization.
image file: d5ra07272g-f6.tif
Fig. 6 Electronic charge density distribution map of Hf2ClC (left) and Hf2BrC (right), respectively.

Significant charge accumulation between Hf and C atoms in both compounds demonstrates strong covalent bonding between these atoms. This strong covalent character arises from the hybridization of Hf-5d and C-2p orbitals, which is consistent with the band structure and density of states (DOS) analysis. The substitution of Cl with Br does not significantly alter this Hf–C bonding, as it remains the dominant interaction in both materials. The Hf–A (A = Cl, Br) bonds also exhibit covalent character but are relatively weaker than Hf–C bonds. The electronic charge density around Cl and Br atoms is more localized, indicating weaker hybridization with Hf compared to Hf–C. The larger atomic radius and lower electronegativity of Br compared to Cl reduce the extent of orbital overlap with Hf, weakening the Hf–Br bond relative to Hf–Cl. This reduced bonding strength is also reflected in the flatter bands in the electronic band structure of Hf2BrC, which suggests lower charge carrier mobility. A noticeable feature in the charge density maps is the ionic nature of Cl–C and Br–C bonds. In both compounds, there is a clear separation of charge between A-site (Cl/Br) and C atoms, confirming the ionic bonding character. However, the charge distribution around Br appears more diffuse compared to Cl, indicating a more polarized bond due to its larger size and lower electronegativity. This increased ionic nature further reduces the overall hybridization in Hf2BrC, leading to a less dispersive electronic structure compared to Hf2ClC.

3.9 Bond population analysis

The Mulliken atomic population analysis provides valuable insights into the nature of chemical bonding and charge transfer in Hf2AC (A = Cl, Br) MAX compounds. The deviation of Mulliken charges from their formal ionic states indicates significant charge redistribution among the atoms, suggesting the coexistence of covalent and ionic bonding. The substitution of Cl with Br affects the charge transfer and bonding characteristics due to differences in electronegativity and atomic size between the two halogen elements.

In Hf2ClC, the Mulliken charge on Cl is −0.26e, while in Hf2BrC, Br exhibits a positive charge of 0.44e. This indicates that Br has a lower tendency to attract electrons compared to Cl, leading to weaker charge transfer in Hf2BrC. The charge distribution also affects the bonding nature; Hf–Cl bonding exhibits stronger hybridization than Hf–Br, as reflected in the calculated effective valence charge (EVC). The EVC values for all species in both compounds are greater than zero, confirming the partial covalency of the bonds.71 However, the effective valence of Hf in Hf2BrC is slightly higher (3.79e) than in Hf2ClC (3.44e), implying that the bonding in Hf2BrC is less ionic and more covalent in nature. This phenomenon is primarily due to the larger atomic radius of Br, which reduces electrostatic attraction and weakens the bond strength between Hf and Br compared to Hf and Cl.

To further validate the bonding characteristics, the Hirshfeld population analysis (HPA) was performed, revealing variations from the Mulliken charge analysis.72 Despite the differences, both methods consistently indicate that the Hf–C bonds remain predominantly covalent, while the Hf–Cl and Hf–Br bonds exhibit partial covalency with some ionic character. The Cl–C and Br–C bonds, however, are primarily ionic in both cases. The substitution of Cl with Br also affects electronic properties, as weaker bonding in Hf2BrC leads to lower charge transfer efficiency and increased carrier localization. This results in reduced electronic conductivity and higher anisotropy in transport properties. However, the relatively weaker Hf–Br bonding may lead to a slight reduction in structural robustness compared to Hf2ClC. Overall, the replacement of Cl with Br alters the bonding nature by decreasing the degree of charge transfer and reducing hybridization effects. This leads to a shift in electronic and transport properties, making Hf2BrC slightly less ionic and more covalent but with weaker bonding interactions compared to Hf2ClC.

3.10 Optical properties

The optical properties of materials are important because they determine their electronic nature and potential applications in optoelectronic devices, solar cells, lasers, photodetectors, sensors, and reconfigurable photonics. The optical behavior of Hf2ClC and Hf2BrC MAX phases is analyzed through the frequency-dependent complex dielectric function, ε(ω) = ε1(ω) + 2(ω), which directly correlates with the electronic structure across all photon energy levels.73 The imaginary part ε2(ω) can be calculated from the momentum matrix elements connecting to the occupied and unoccupied electronic states with the following relations:74
 
image file: d5ra07272g-t28.tif(24)
where, the integral is over the first Brillouin zone, ω refers the frequency of light, e indicates the electronic charge, u is the vector indicating the polarization of the incident electric field, ε0 is the dielectric constant in free space and Ψkc and Ψkv are the conduction and valence band wave functions respectively. The real part ε1(ω) was obtained from the imaginary part ε2(ω) by using the Kramer–Kronig (K–K) relations:73
 
image file: d5ra07272g-t29.tif(25)
where P implies the principal value of the integral. In order to describe the material response to incident electromagnetic radiation, for studied MAX phase compounds Hf2AC (A = Cl, Br) all the other energy-dependent optical parameters, like absorption coefficient α(ω), photoconductivity σ(ω), loss function L(ω), reflectivity R(ω) and refractive index, n(ω) were computed directly from ε1(ω) and ε2(ω) and displayed in Fig. 7, respectively. These properties were examined across a photon energy range of up to 30 eV. As in the case of many MAX phases, Hf2AC (A = Cl, Br) exhibits metallic behavior, which significantly influences their optical properties. The conduction electrons in these materials affect the dielectric function by changing how they respond to light. Specifically, the intra-band contributions from free electrons play a dominant role, affecting the low-energy optical spectrum. In metallic or metal-like conductors, both intra-band and inter-band transitions shape the optical behavior, requiring a precise approach for accurate characterization. To achieve a reliable analysis of the optical properties of Hf2ClC and Hf2BrC, the conventional Drude model was employed.75 This model effectively accounts for free-electron contributions to the dielectric function. For our calculations, we incorporated a Drude damping parameter of 0.05 eV and a plasma frequency of 3 eV for both compounds. These values are consistent with earlier reports for metallic MAX phases and related carbides/nitrides,76–80 where such parameters accurately describe the free-electron contributions and Drude-like behavior at low photon energies. The Drude term primarily captures the intra-band transitions, which dominate at low photon energies. These transitions arise due to the free-electron nature of the material and contribute significantly to optical conductivity, reflectivity, and absorption spectra.

image file: d5ra07272g-f7.tif
Fig. 7 The energy dependent (a) real part of dielectric function, (b) imaginary part of dielectric function, (c) refractive index, (d) extinction coefficient (e) absorption coefficient, (f) photoconductivity, (g) reflectivity, and (h) loss function of Hf2ClC and Hf2BrC, respectively.
3.10.1 Dielectric function. The optical behavior of Hf2ClC and Hf2BrC provides a compelling insight into their metallic nature, as depicted by the real ε1(ω) and imaginary ε2(ω) components of their dielectric function, which are depicted in Fig. 7(a and b). The presence of Cl and Br in these MAX phases leads to subtle yet significant differences in their optical responses. The real part ε1(ω) exhibits a strong metallic character, showing negative values at low energy, which is consistent with the Drude-like behavior typically observed in conductive materials. This trend aligns well with the electronic band structure, reaffirming the metallic conductivity of these compounds. As shown in Fig. 7(a), ε1(ω) attains large negative values at low photon energies and crosses zero at approximately 21.79 eV for Hf2ClC and 20.77 eV for Hf2BrC. This zero-crossing energy corresponds to the plasma frequency, where the absorption coefficient (Fig. 7(e)) vanishes, reflectivity (Fig. 7(g)) drops sharply, and the energy loss function (Fig. 7(h)) exhibits its first pronounced peak. Moreover, as depicted in Fig. 7(a) and (b), ε1(ω) crosses zero from below and approaches zero from above, thereby reinforcing the metallic and highly reflective characteristics.

The imaginary part of the dielectric function, ε2(ω), which represents optical absorption due to electron excitations, is shown in Fig. 7(b). The high ε2(ω) values at low photon energies indicate strong absorption for both Hf2ClC and Hf2BrC. As seen in Fig. 7(b), ε2(ω) initially drops to ε2 = 15.59 at 0.57 eV for Hf2ClC and to ε2 = 12.69 at 0.56 eV for Hf2BrC, before rising to their respective maxima of 16.84 at 0.97 eV and 15.65 at 1.13 eV. Beyond these peaks, ε2(ω) gradually decreases almost exponentially across 1.38–6.49 eV and 8.24–15.77 eV for Hf2ClC, and 3.35–7.93 eV for Hf2BrC, reflecting their optical stability against excessive energy losses. The low-energy peaks originate from intra-band transitions of conduction electrons, while the high-energy peaks are attributed to inter-band transitions between occupied and unoccupied states, consistent with metallic characteristics. The high static dielectric constant of ε2(ω) at low photon energies also signifies strong light–matter interactions, highlighting the potential of Hf2ClC and Hf2BrC for high-k dielectric, capacitor, and optical modulator applications.

3.10.2 Refractive index. The refractive index, n(ω), is a key optical parameter that governs how light propagates through a material, making it fundamental for photonic and optoelectronic applications. The variation of n(ω) with photon energy for Hf2ClC and Hf2BrC, shown in Fig. 7(c), exhibits a gradual decrease with increasing energy. Both compounds display remarkably high static refractive index values of n(0) = 84.33 for Hf2ClC and 84.31 for Hf2BrC, indicating strong light–matter interactions and high polarizability, which are advantageous for waveguides, photonic crystals, and energy-harvesting devices. The refractive index decreases sharply with photon energy, showing prominent peaks near 1.10 eV (n = 3.58) for both phases. Within the visible range (1.67–3.10 eV), n(ω) declines from 3.06 to 2.08 for Hf2ClC and from 3.18 to 2.26 for Hf2BrC. The slightly higher refractive index in Hf2BrC can be attributed to Br's larger atomic radius and lower electronegativity, which modify the electronic structure and enhance polarization effects. This tunable refractive response, combined with metallic conductivity, demonstrates the advantages of these MAX phases for advanced optoelectronic and photonic devices requiring efficient light modulation and energy management.
3.10.3 Extinction coefficient. The extinction coefficient, k(ω), a key optical parameter that quantifies the attenuation of electromagnetic radiation as it propagates through a material, reflects intrinsic absorption losses. This property provides valuable insight into light–matter interactions and is particularly significant for photonics, optical coatings, and optoelectronic device applications. The calculated k(ω) spectra for Hf2ClC and Hf2BrC, presented in Fig. 7(d), exhibit distinct frequency-dependent trends. As shown in Fig. 7(d), both compounds display high static values, k(0) ≈ 102.64 for Hf2ClC and 102.66 for Hf2BrC, with strong absorption features in the infrared region. With increasing photon energy, k(ω) gradually decreases; however, Hf2BrC reaches a slightly higher peak (k = 2.11 at 8.04 eV) compared to Hf2ClC (k = 1.88 at 8.55 eV) in the visible range, indicating enhanced photon–electron coupling due to Br substitution. The pronounced extinction values signify efficient light absorption and strong photon-induced charge excitation, desirable for optical and photoactive devices. Additionally, k(ω) exhibits minor anisotropy within the 3.52–20.13 eV range, consistent with the layered structure of MAX phases. The observed local maxima in the k(ω) spectra coincide with the zero-crossing points of the real part of the dielectric function, confirming the correlation between energy loss and plasma resonance behavior.
3.10.4 Absorption coefficient. The absorption coefficient, α(ω), provides vital information about a material's ability to absorb incident photon energy, directly influencing its optoelectronic performance. The absorption spectra of Hf2ClC and Hf2BrC, illustrated in Fig. 7(e), reveal strong photon–matter interactions extending across the infrared (IR), visible, and ultraviolet (UV) regions. Due to their metallic character, absorption begins at 0 eV and exhibits primary peaks at 0.18 eV (α = 0.46 × 105 cm−1), followed by secondary peaks in the visible range near 2.69 eV, where α reaches approximately 0.84 × 105 cm−1 for both phases. Maximum absorption occurs in the UV region at 12.15 eV (α = 2.84 × 105 cm−1) for Hf2ClC and 14.79 eV (α = 2.91 × 105 cm−1) for Hf2BrC. Within the visible spectrum (1.67–3.10 eV), α(ω) increases steadily from 0.61 × 105 cm−1 to 0.83 × 105 cm−1 for Hf2ClC and from 0.59 × 105 cm−1 to 0.81 × 105 cm−1 for Hf2BrC, indicating excellent light-harvesting efficiency. This consistent absorption behavior highlights the strong electronic transition activity within the conduction band, driven primarily by Hf-d orbital contributions near the Fermi level. The broad and intense absorption response across a wide energy range underscores the potential of these MAX phases for high-performance photodetectors, UV-visible absorbers, and other optoelectronic and photonic devices.
3.10.5 Photoconductivity. The photoconductivity spectra of Hf2ClC and Hf2BrC, illustrated in Fig. 7(f), commence at zero photon energy, confirming the absence of a band gap and thereby validating their metallic nature. This behavior arises from the presence of free charge carriers that contribute to electrical conductivity even at low photon energies. The observed trends in photoconductivity are consistent with the absorption coefficient, electronic band structure, and density of states (DOS) analyses, all of which indicate substantial carrier mobility and metallic behavior. The frequency-dependent optical conductivity, σ(ω), of Hf2AC (A = Cl, Br) further highlights strong photon–electron interactions essential for charge transport processes. The static optical conductivity of these compounds is about 20.93 S m−1, which means that they can move charges around easily and excite electrons quickly when light hits them. Within the visible region, σ(ω) varies between 2.77–2.59 S m−1 for Hf2ClC and 2.82–2.75 S m−1 for Hf2BrC, with the latter exhibiting slightly enhanced conductivity. The maximum photoconductivity is observed at 8.0 eV for Hf2ClC (σ = 6.31 S m−1) and at 7.67 eV for Hf2BrC (σ = 7.18 S m−1), after which σ(ω) decreases gradually with energy. The combination of strong low-energy conductivity and photon-induced carrier excitation underscores the potential of these MAX phases for optoelectronic and photoactive applications, including photodetectors, solar energy converters, and optical sensors, where efficient light-to-charge conversion is vital.
3.10.6 Reflectivity. The reflectivity spectra of Hf2ClC and Hf2BrC, presented in Fig. 7(g), exhibit similar overall trends, characteristic of metallic materials. Both compounds have high static reflectivity values of R(0) ≈ 0.98, which shows that they are very good at reflecting low-energy electromagnetic radiation. As photon energy increases, the reflectivity decreases, reaching R = 0.46 at 0.82 eV for Hf2ClC and R = 0.43 at 0.75 eV for Hf2BrC. The spectra of both compounds follow a comparable decreasing trend up to approximately 3.44 eV, beyond which a slight anisotropy is observed, reflecting directional dependence in their optical response due to their layered crystal structure. Within the visible photon energy range (1.67–3.10 eV), Hf2ClC and Hf2BrC maintain moderate reflectivity levels between 43% and 32%, indicating controlled light reflection suitable for solar radiation management and optical coating applications. The high reflectivity in the low-energy region coupled with tunable optical anisotropy at higher energies suggests that these MAX phases could serve as efficient reflective coatings and thermal control materials for optoelectronic and high-temperature devices.
3.10.7 Loss function. The energy loss function, L(ω), represents the energy dissipated by fast-moving electrons traversing a material and provides critical insight into its plasmonic and dielectric behavior. As illustrated in Fig. 7(h), the L(ω) spectra for Hf2ClC and Hf2BrC exhibit no distinct peaks in the low-energy range (0–15 eV), which can be attributed to the relatively large values of the imaginary component of the dielectric function, ε2(ω), that dampen collective oscillations. A prominent peak emerges between 15 and 28 eV for both compounds, corresponding to bulk plasma oscillations—collective excitations of conduction electrons within the material. The maximum peak in each spectrum indicates the bulk plasma frequency, where the real part of the dielectric function crosses zero, marking the transition from metallic to dielectric behavior. The calculated plasma frequencies are 22.21 eV for Hf2ClC and 22.35 eV for Hf2BrC, indicating a slightly stronger plasmonic response in the Br-substituted phase. These energy-loss peaks coincide with sharp decreases in reflectivity, signifying the point at which the material transitions from a reflective to a transparent regime. Beyond the plasma frequency, both compounds exhibit minimal energy loss and increased transparency, suggesting their potential for high-frequency optical and plasmonic applications where low optical damping and efficient charge oscillations are essential.

4. Conclusions

This first-principles study reveals that A-site halogen substitution is an effective approach for tailoring the multifunctional behavior of Hf2AC (A = Cl, Br) MAX phases. Both Hf2ClC and Hf2BrC phases are confirmed to be structurally, mechanically, and dynamically stable. The substitution of Cl with Br reduces stiffness, hardness, and sound velocity, enhancing ductility, vibration damping, and thermal insulation performance, suitable for thermal barrier applications. In contrast, Hf2ClC exhibits greater rigidity, higher conductivity, and superior thermal dissipation, making it better suited for structural and heat-management uses. Electronically, both compounds display metallic conductivity dominated by Hf–C bonding, while optically, Br substitution shifts the absorption toward lower energies, improving infrared and UV responsiveness. Overall, halogen engineering at the A-site emerges as a powerful strategy for tuning mechanical, thermal, and optical functionalities in Hf-based MAX phases for diverse high-temperature and optoelectronic applications.

Author contributions

Asif Mohammed Arfi: investigation, methodology, data curation, conceptualization, writing original manuscript; Mahamudun Nabi: writing – original draft, methodology, formal analysis, data calculations, validation. Md. Hasan Mia: writing – original draft, review & editing, conceptualization, formal analysis, validation, supervision. Omar Alsalmi: data curation, formal analysis, review – editing; Muhammad Athar Uddin data curation, formal analysis, review – editing; Md. Zahid Hasan: formal analysis, validation, supervision, review – editing.

Conflicts of interest

The authors declare that they have no known conflicting financial interests or personal ties that may have seemed to affect the work presented in this study.

Data availability

Relevant data from this study are available from the corresponding author upon a reasonable request.

Acknowledgements

The authors extend their appreciation to Umm Al-Qura University, Saudi Arabia for funding this research work through grant number: 25UQU4300099GSSR10. Funding statement: This research work was funded by Umm Al-Qura University, Saudi Arabia under grant number: 25UQU4300099GSSR10.

References

  1. M. A. Hadi, Superconducting phases in a remarkable class of metallic ceramics, J. Phys. Chem. Solid., 2020, 138, 109275 CrossRef CAS.
  2. D. Ohmer, Q. Gao, I. Opahle, H. K. Singh and H. Zhang, High-throughput design of 211-M 2 AX compounds, Phys. Rev. Mater., 2019, 3(5), 053803 CrossRef CAS.
  3. I. Ouadha, R. Habib, A.-R. Ahmed, A. Reggad and D. Rached, Study of the structural, mechanical and thermodynamic properties of the new MAX phase compounds (Zr1-xTix) 3AlC2, Comput. Condens. Matter, 2020, 23, e00468 CrossRef.
  4. A. Rached, H. R. Ahmed, M. Haque Babu, T. Hadji and D. Rached, Prediction of double transition metal (Cr1−xZrx) 2AlC MAX phases as thermal barrier coatings: Insight from density functional theory, Int. J. Quant. Chem., 2021, 121(20), e26770 CrossRef.
  5. M. H. Mia, U. Ahmed, S. K. Saha and Md A. Ali, Unveiling mechanical, electronic, and optical properties of newly synthesized Mo2VAlC2 and Mo2V2AlC3 o-MAX phases via first-principles calculations, J. Mater. Res. Technol., 2025, 36, 2468–2484 CrossRef CAS.
  6. Y. Katoh, L. L. Snead, T. Cheng, C. Shih, W. Daniel Lewis, T. Koyanagi, T. Hinoki, C. H. Henager Jr and F. Monica, Radiation-tolerant joining technologies for silicon carbide ceramics and composites, J. Nucl. Mater., 2014, 448(1–3), 497–511 CrossRef CAS.
  7. M. A. Khatun, M. H. Mia, S. Ahmad and S. K. Mitro, Crystal stability and the origin of transport properties of new MAX phases Y2AN (A= In, Tl): a theoretical background for experimental study, Heliyon, 2025, e42646 CrossRef CAS.
  8. T. Cabioch, P. Eklund, M. Vincent, M. Jaouen and M. W. Barsoum, Tailoring of the thermal expansion of Cr2 (Alx, Ge1−x) C phases, J. Eur. Ceram. Soc., 2013, 33(4), 897–904 CrossRef CAS.
  9. M. Radovic, M. W. Barsoum, A. Ganguly, T. Zhen, P. Finkel, S. R. Kalidindi and E. Lara-Curzio, On the elastic properties and mechanical damping of Ti3SiC2, Ti3GeC2, Ti3Si0. 5Al0. 5C2 and Ti2AlC in the 300–1573 K temperature range, Acta Mater., 2006, 54(10), 2757–2767 CrossRef CAS.
  10. J. C. Schuster, H. Nowotny and C. Vaccaro, The ternary systems: CrAlC, VAlC, and TiAlC and the behavior of H-phases (M2AlC), J. Solid State Chem., 1980, 32(2), 213–219 CrossRef CAS.
  11. M. W. Barsoum, I. Salama, T. El-Raghy, J. Golczewski, H. J. Seifert, F. Aldinger, W. D. Porter and H. Wang, Thermal and electrical properties of Nb 2 AlC, (Ti, Nb) 2 AlC and Ti 2 AlC, Metall. Mater. Trans. A, 2002, 33, 2775–2779 CrossRef.
  12. I. Salama, T. El-Raghy and M. W. Barsoum, Synthesis and mechanical properties of Nb2AlC and (Ti, Nb) 2AlC, J. Alloys Compd., 2002, 347(1–2), 271–278 CrossRef.
  13. S. Gupta and M. W. Barsoum, Synthesis and oxidation of V 2AlC and (Ti0. 5, V 0.5) 2AlC in air, J. Electrochem. Soc., 2004, 151(2), D24 CrossRef CAS.
  14. H. Yang, B. Manoun, R. T. Downs, A. Ganguly and M. W. Barsoum, Crystal chemistry of layered carbide, Ti3 (Si0. 43Ge0. 57) C2, J. Phys. Chem. Solids, 2006, 67(12), 2512–2516 CrossRef CAS.
  15. B. Manoun, S. K. Saxena, G. Hug, A. Ganguly, E. N. Hoffman and M. W. Barsoum, Synthesis and compressibility of Ti3 (Al, Sn0. 2) C2 and Ti3Al (C0. 5, N0. 5) 2, J. Appl. Phys., 2007, 101, 11 CrossRef.
  16. M. W. Barsoum and M. Radovic, ‘Elastic and mechanical properties of the MAX phases, Annu. Rev. Mater. Res., 2011, 41(1), 195–227 CrossRef CAS.
  17. H. Kim, J. Choi, J. Choi, J. Kwak and J. Kim, Ab initio simulations of M2AX (M= lanthanides, A= Al or Si, X= C or N) phases as bond coats for thermal barrier coatings, Scr. Mater., 2025, 258, 116520 CrossRef CAS.
  18. O. S. Rijal, H. Krishna Neupane, D. Oli, R. Kiran Neupane, P. Shrestha, S. Sharma, L. P. Joshi and P. Rajendra, A first-principles investigation of the structural, mechanical, dynamic, electronic, magnetic, and optical properties of Ti2AC (A= Cd, S) MAX phase compounds, J. Phys. D: Appl. Phys., 2025, 58(12), 125102 CrossRef CAS.
  19. B. U. Haq, Se-H. Kim, R. Ahmed, A. Rasool Chaudhry, M. Shabbir and A. Laref, Physical properties of W2GaX (X= C, N, and F) based novel MAX phases; potential materials for applications in advanced electronic and optical devices, Comput. Theor. Chem., 2024, 1231, 114427 CrossRef.
  20. M. N. Amin, A. A. Faysal, A. Azzouz-Rached, F. I. Chowdhury, J. Uddin, J. Uddin, J. Islam, A. R. Piyal, A. K. Chowdhury and A. K. M. Rahman, DFT insights into the structural, mechanical, electronic, optical, and thermal properties of Cr2AC (A= Si, Al, Ga, Ge, and P) MAX phases carbides, AIP Adv., 2024, 14(12), 125302 CrossRef CAS.
  21. W. Tan, Y. Tian, Y. Zhou, X. Wei, L. Zhang, X. Tao and Y. Ouyang, Trends in mechanical, anisotropic, electronic, and thermal properties of MAX phases: a DFT study on M2SX phases, Mater. Today Commun., 2023, 35, 105759 CrossRef CAS.
  22. M. A. Ali and W. Q. Muhammad, DFT insights into the new Hf-based chalcogenide MAX phase Hf2SeC, Vacuum, 2022, 201, 111072 CrossRef CAS.
  23. M. M. Uddin, MdA. Ali, MdM. Hossain, S. H. Naqib and A. K. M. A. Islam, Comparative study of predicted MAX phase Hf 2 AlN with recently synthesized Hf 2 AlC: a first principle calculations, Indian J. Phys., 2022, 1–13 CAS.
  24. W. Kohn and J. S. Lu, Self-consistent equations including exchange and correlation effects, Phys. Rev., 1965, 140(4A), A1133 CrossRef.
  25. S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. I. J. Probert, K. Refson and M. C. Payne, First principles methods using CASTEP, Z. Kristallogr. Cryst. Mater., 2005, 220(5–6), 567–570 CrossRef CAS.
  26. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77(18), 3865 CrossRef CAS PubMed.
  27. H. J. Monkhorst and J. D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B, 1976, 13(12), 5188 CrossRef.
  28. D. Vanderbilt, Soft self-consistent pseudopotentials in a generalized eigenvalue formalism, Phys. Rev. B:Condens. Matter Mater. Phys., 1990, 41(11), 7892 CrossRef PubMed.
  29. T. H. Fischer and A. Jan, General methods for geometry and wave function optimization, J. Phys. Chem., 1992, 96(24), 9768–9774 CrossRef CAS.
  30. C. M. Brown, W. Dreyer and W. H. Müller, The convergence of a DFT-algorithm for solution of stress-strain problems in composite mechanics, J. Eng. Mater. Technol., 2003, 125(1), 27–37 CrossRef.
  31. K. Parlinski, Z. Q. Li and Y. Kawazoe, First-principles determination of the soft mode in cubic ZrO2, Phys. Rev. Lett., 1997, 78(21), 4063 CrossRef CAS.
  32. K. Momma and F. Izumi, VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data, Appl. Crystallogr., 2011, 44(6), 1272–1276 CrossRef CAS.
  33. A. Bouhemadou, Structural, Electronic and Elastic Properties of Ti 2 TlC, Zr 2 TlC and Hf 2 TlC, Cent. Eur. J. Phys., 2009, 7, 753–761 CAS.
  34. K. Akter, F. Parvin, M. A. Hadi and A. K. M. A. Islam, Insights into the predicted Hf2SN in comparison with the synthesized MAX phase Hf2SC: A comprehensive study, Comput. Condens. Matter, 2020, 24, e00485 CrossRef.
  35. A. Roumili, Y. Medkour and D. Maouch, Elastic and Electronic Properties of Hf 2 SnC and Hf 2 SnN, Int. J. Mod. Phys. B, 2009, 23(26), 5155–5161 CrossRef CAS.
  36. T. Lapauw, B. Tunca, T. Cabioc'h, J. Lu, P. O. Å. Persson, K. Lambrinou and J. Vleugels, Synthesis of MAX phases in the Hf–Al–C system, Inorg. Chem., 2016, 55(21), 10922–10927 CrossRef CAS PubMed.
  37. Q. Zhang, Y. Zhou, X. San, W. Li, Y. Bao, Q. Feng, S. Grasso and C. Hu, Zr2SeB and Hf2SeB: Two new MAB phase compounds with the Cr2AlC-type MAX phase (211 phase) crystal structures, J. Adv. Ceram., 2022, 11(11), 1764–1776 CrossRef CAS.
  38. M. B. Kanoun, S. Goumri-Said and M. Jaouen, Steric effect on the M site of nanolaminate compoundsM2SnC (M= Ti, Zr, Hf and Nb), J. Phys.:Condens. Matter, 2009, 21(4), 045404 CrossRef CAS.
  39. M. H. Mia and A. K. Mst, Dynamic tuning of optoelectronic and mechanical properties in TlMCl3 (M= Ge, Sn) under pressure-induced phase transition, Heliyon, 2025, e42603 CrossRef CAS PubMed.
  40. A. B. Alruqi, Engineering the Mechanics and Thermodynamics of Ti3AlC2, Hf3AlC2, Hf3GaC2,(ZrHf) 3AlC2, and (ZrHf) 4AlN3 MAX Phases via the Ab Initio Method, Crystals, 2025, 15(1), 87 CrossRef CAS.
  41. M. Dahlqvist, B. Alling and J. Rosén, Stability trends of MAX phases from first principles, Phys. Rev. B:Condens. Matter Mater. Phys., 2010, 81(22), 220102 CrossRef.
  42. L. Sun, L. Miao and Y. Cao, The physical properties of new MAX phase Zr3XC2 (X= Cd, Sb): A DFT investigation, J. Mater. Sci., 2025, 1–19 Search PubMed.
  43. S. Kirklin, J. E. Saal, B. Meredig, A. Thompson, J. W. Doak, M. Aykol, S. Rühl and C. Wolverton, The Open Quantum Materials Database (OQMD): assessing the accuracy of DFT formation energies, npj Comput. Mater., 2015, 1(1), 1–15 Search PubMed.
  44. N. Goossens, B. Tunca, K. Lambrinou and J. Vleugels, Topotactic transformations in Hf-Al-C MAX phase compounds: Synthesis and characterization of nanolaminated Hf2AlC, Hf3AlC2 and Hf5Al2C3, Open Ceram., 2024, 18, 100584 CrossRef CAS.
  45. L. Chen, Y. Li, K. Chen, X. Bai, M. Li, S. Du, Z. Chai and Q. Huang, Synthesis and characterization of medium-/high-entropy M2SnC (M= Ti/V/Nb/Zr/Hf) MAX phases, Small Struct., 2023, 4(1), 2200161 CrossRef CAS.
  46. Q. Zhang, B. Wen, L. Jia, Y. Zhou, X. San, Y. Bao, L. Chu, Q. Feng, S. Grasso and C. Hu, Synthesis of new lead-containing MAX phases of Zr3PbC2 and Hf3PbC2, J. Am. Ceram. Soc., 2023, 106(11), 6390–6397 CrossRef CAS.
  47. F. Mouhat and F.-X. Coudert, Necessary and sufficient elastic stability conditions in various crystal systems, Phys. Rev. B:Condens. Matter Mater. Phys., 2014, 90(22), 224104 CrossRef.
  48. M. I. Naher and S. H. Naqib, An ab-initio study on structural, elastic, electronic, bonding, thermal, and optical properties of topological Weyl semimetal Ta X (X= P, As), Sci. Rep., 2021, 11(1), 5592 CrossRef CAS.
  49. M. H. Mia, F. Parvin, A. K. M. A. Islam and M. A. Khatun, Ternary chalcogenides NbIn X 2 (X= S, Se): A comprehensive investigation of mechanical, electronic, vibrational, optical and thermophysical properties, Int. J. Mod. Phys. B, 2025, 39(14), 2550118 CrossRef CAS.
  50. R. Hill, The elastic behaviour of a crystalline aggregate, Proc. Phys. Soc., London, Sect. A, 1952, 65(5), 349 CrossRef.
  51. Y. W. Bao, X. H. Wang, H. B. Zhang and Y. C. Zhou, Thermal shock behavior of Ti3AlC2 from between 200 °C and 1300 °C, J. Eur. Ceram. Soc., 2005, 25(14), 3367–3374 CrossRef CAS.
  52. M. A. Khatun, M. H. Mia, M. A. Hossain, F. Parvin and A. K. M. A. Islam, Optical and thermoelectric properties of layer structured Ba2XS4 (X= Zr, Hf) for energy harvesting applications, J. Phys. Chem. Solids, 2025, 196, 112381 CrossRef CAS.
  53. P. H. Mott, J. R. Dorgan and C. M. Roland, The bulk modulus and Poisson's ratio of “incompressible” materials, J. Sound Vib., 2008, 312(4–5), 572–575 CrossRef.
  54. X.-Q. Chen, H. Niu, D. Li and Y. Li, Modeling hardness of polycrystalline materials and bulk metallic glasses, Intermetallics, 2011, 19(9), 1275–1281 CrossRef CAS.
  55. N. Miao, S. Baisheng, J. Zhou and Z. Sun, Theoretical investigation on the transition-metal borides with Ta3B4-type structure: A class of hard and refractory materials, Comput. Mater. Sci., 2011, 50(4), 1559–1566 CrossRef CAS.
  56. Md A. Rahman, M. Asma Khatun, Md H. Mia, M. Zugrufa and I. Ahmad, First-Principles Calculations to Investigate the Structural, Mechanical, Electronic, Optical and Thermal Properties of Disilicides Materials UT2Si2 (T= Ru, Rh, and Pd), Phys. Status Solidi B, 2025, 262(1), 2400230 CrossRef CAS.
  57. P. Ravindran, L. Fast, P. A. Korzhavyi, B. Johansson, J. Wills and E. Os, Density functional theory for calculation of elastic properties of orthorhombic crystals: Application to TiSi 2, J. Appl. Phys., 1998, 84(9), 4891–4904 CrossRef CAS.
  58. C. M. Kube and D. J. Maarten, Elastic constants of polycrystals with generally anisotropic crystals, J. Appl. Phys., 2016, 120(16), 165105 CrossRef.
  59. S. I. Ranganathan and M. Ostoja-Starzewski, Universal elastic anisotropy index, Phys. Rev. Lett., 2008, 101(5), 055504 CrossRef PubMed.
  60. M. H. Mia, M. A. Khatun and M. Rahman, Ti2PTe2 chalcogenide: A comprehensive DFT study on physical properties, Next Mater., 2025, 7, 100434 CrossRef.
  61. E. Schreiber, O. L. Anderson, N. Soga, and J. F. Bell, ‘Elastic Constants and Their Measurement’, 1975, pp. 747–748 Search PubMed.
  62. O. L. Anderson, A simplified method for calculating the Debye temperature from elastic constants, J. Phys. Chem. Solid., 1963, 24(7), 909–917 CrossRef CAS.
  63. M. E. Fine, L. D. Brown and H. L. Marcus, Elastic constants versus melting temperature in metals, Scripta Metall., 1984, 18(9), 951–956 CrossRef CAS.
  64. V. N. Belomestnykh and E. P. Tesleva, Interrelation between Anharmonicity and Lateral Strain in Quasi-Isotropic Polycrystalline Solids, Tech. Phys., 2004, 49(8), 1098 CrossRef CAS.
  65. C. Dhakal, S. Aryal, R. Sakidja and W.-Y. Ching, Approximate lattice thermal conductivity of MAX phases at high temperature, J. Eur. Ceram. Soc., 2015, 35(12), 3203–3212 CrossRef CAS.
  66. C. L. Julian, Theory of heat conduction in rare-gas crystals, Phys. Rev., 1965, 137(1A), A128 CrossRef.
  67. D. R. Clarke, Materials selection guidelines for low thermal conductivity thermal barrier coatings, Surf. Coating. Technol., 2003, 163, 67–74 CrossRef.
  68. M. W. Qureshi, X. Ma, G. Tang and P. Ramesh, Structural stability, electronic, mechanical, phonon, and thermodynamic properties of the M2GaC (M= Zr, Hf) max phase: an AB initio calculation, Materials, 2020, 13(22), 5148 CrossRef CAS PubMed.
  69. M. A. Hadi, N. Kelaidis, S. H. Naqib, A. K. M. A. Islam, A. Chroneos and R. V. Vovk, Insights into the physical properties of a new 211 MAX phase Nb2CuC, J. Phys. Chem. Solids, 2021, 149, 109759 CrossRef CAS.
  70. Y. Mo, R. Paul and W. Y. Ching, Electronic structure and optical conductivities of 20 MAX-phase compounds, Phys. Rev. B:Condens. Matter Mater. Phys., 2012, 86(16), 165122 CrossRef.
  71. M. D. Segall, R. Shah, C. J. Pickard and M. C. Payne, Population analysis of plane-wave electronic structure calculations of bulk materials, Phys. Rev. B, 1996, 54(23), 16317 CrossRef CAS PubMed.
  72. F. L. Hirshfeld, Bonded-atom fragments for describing molecular charge densities, Theor. Chim. Acta, 1977, 44, 129–138 CrossRef CAS.
  73. M. Tarekuzzaman, N. Shahadath, M. Montasir, O. Alsalmi, M. H. Mia, H. Al-Dmour, M. Rasheduzzaman and M. Z. Hasan, DFT analysis of the physical properties of direct band gap semiconducting double perovskites A 2 BIrCl 6 (A= Cs, Rb; B= Na, K) for solar cells and optoelectronic applications, RSC Adv., 2025, 15(17), 13643–13661 RSC.
  74. O. Das, J. K. Rony, P. Kumar Karmaker, M. A. Khatun, Md Murshidul Islam, M. H. Mia, M. Saiduzzaman and M. Islam, Tailoring the physical properties of InSnX3 (X= F, cl) perovskites via pressure: A path toward sustainable optoelectronics, Chem. Phys., 2025, 112635 CrossRef CAS.
  75. S. Li, R. Ahuja, M. W. Barsoum, P. Jena and B. Johansson, Optical properties of Ti3SiC2 and Ti4AlN3, Appl. Phys. Lett., 2008, 92, 22 Search PubMed.
  76. M. A. Ali, M. M. Hossain, M. M. Uddin, A. K. M. A. Islam and S. H. Naqib, Understanding the improvement of thermo-mechanical and optical properties of 212 MAX phase borides Zr2AB2 (A= In, Tl), J. Mater. Res. Technol., 2021, 15, 2227–2241 CrossRef CAS.
  77. M. R. Khatun, M. A. Ali, F. Parvin and A. K. M. A. Islam, Elastic, thermodynamic and optical behavior of V2AC (A= Al, Ga) MAX phases, Results Phys., 2017, 7, 3634–3639 CrossRef.
  78. M. W. Qureshi, M. A. Ali and X. Ma, Screen the thermomechanical and optical properties of the new ductile 314 MAX phase boride Zr3CdB4: A DFT insight, J. Alloys Compd., 2021, 877, 160248 CrossRef CAS.
  79. M. A. Hadi, Y. Panayiotatos and A. Chroneos, Structural and optical properties of the recently synthesized (Zr3−x Ti x) AlC2 MAX phases, J. Mater. Sci.: Mater. Electron., 2017, 28(4), 3386–3393 CrossRef CAS.
  80. A. Chowdhury, M. A. Ali, M. M. Hossain, M. M. Uddin, S. H. Naqib and A. K. M. A. Islam, Predicted MAX phase Sc2InC: dynamical stability, vibrational and optical properties, Phys. Status Solidi B, 2018, 255(3), 1700235 CrossRef.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.