Open Access Article
Asif Mohammed Arfi
ae,
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
First published on 5th November 2025
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.
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.
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).
| 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 |
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
![]() | (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
![]() | (2) |
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.
![]() | ||
| 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.,
). 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) |
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 = (C11 − C12)/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) |
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′′ = (C12 − C44) 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
, 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).
| 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 |
![]() | ||
| 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
, 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
![]() | (5) |
| Hmacro = 2(k2G)0.585 − 3 | (6) |
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.
![]() | (7) |
![]() | (8) |
![]() | (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.
| 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 = (BV − BR)/(BV + BR) | (10) |
| AG = (GV − GR) + (GV + GR) | (11) |
![]() | (12) |
![]() | (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
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
![]() | (14) |
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.
![]() | (15) |
![]() | (16) |
![]() | (17) |
Acoustic impedance (Z) of a medium is another important parameter that can be calculated from its density (ρ) and shear modulus (G):49
![]() | (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
![]() | (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.
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.
![]() | (20) |
| 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
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
![]() | (21) |
![]() | (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
![]() | (23) |
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
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
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
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.
![]() | ||
| Fig. 4 Calculated band structure of (a) Hf2ClC and (b) Hf2BrC along the high symmetry directions in the Brillouin zone at ambient conditions. | ||
| 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.
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.
![]() | ||
| 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.
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.
![]() | (24) |
![]() | (25) |
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.
| This journal is © The Royal Society of Chemistry 2025 |